Skip to main content

Full text of "The journal of raptor research"

See other formats


The Journal of 



Volume 39 Number 4 December 2005 






THE RAPTOR RESEARCH FOUNDATION, INC. 

(Founded 1966) 

http: //biology.boisestate.edu/ raptor/ 

OFFICERS 

SECRETARY; Judith Henckel 
TREASURER; Jim Fitzpatrick 

BOARD OF DIRECTORS 

INTERNATIONAL DIRECTOR #3; 

Steve Redpath 

DIRECTOR AT LARGE #1; Jemima ParryJones 
DIRECTOR AT LARGE #2: Eduardo Inigo-Euas 
DIRECTOR AT LARGE #3; Michael W. Collopy 
DIRECTOR AT LARGE #4; Carol McIntyre 
DIRECTOR AT LARGE #5; John A Smallwood 
DIRECTOR AT LARGE #6; Daniel E. Varland 

Ruth Tingay 


NORTH AMERIGAN DIRECTOR #1: 
Steve Hoffman 

NORTH AMERICAN DIRECTOR #2: 
Gary Santolo 

NORTH AMERICAN DIRECTOR #3: 
Ted Swem 

INTERNATIONAL DIRECTOR#!: 
Nick Mooney 

INTERNATIONAI. DIRECTOR #2: 


PRESIDENT; Brian A. Millsap 
VICE-PRESIDENT; David M. Bird 


^ 

EDITORIAL STAEE 


EDITOR; James C. Bednarz, Department of Biological Sciences, RO. Box 599, Arkansas State 
University, State University, AR 72467 U.S.A. 


ASSOCIATE EDITORS 

James R. Beltpioff Joan L. Morrison 

Clint W. Boal Eabrizio Sergio 

Cheryl R. Dykstra Ian G. Warkentin 

Michael I. Goldstein James W. Watson 

BOOK REVIEW EDITOR; Joelle Gehring, Department of Biology, Central Michigan University, 

Mount Pleasant, MI 48859 U.S.A. 

SPANISH EDITORS; Carlos Daniel Cadena, Lucio R. Malizia, Cintia Cornelius 
EDITORIAL ASSISTANTS: Autumn Farless, Joan Ciark 

The Journal of Raptor Research is distributed quarterly to all current members. Original manuscripts 
dealing with the biology and conservation of diurnal and nocturnal birds of prey are welcomed from 
throughout the world, but must be written in English. Submissions can be in the form of research articles, 
short communications, letters to the editor, and book reviews. Contributors should submit either electron- 
ic tiles or a typewritten original and three copies to the Editor. All submissions must be typewritten and dou- 
ble-spaced on one side of 216 X 278 mm (8% XII in.) or standard international, white, bond paper, with 
25 mm (1 in.) margins. The cover page should contain a tide, the author’s full name(s) and address(es). 
Name and address should be centered on the cover page. If the current address is different, indicate this 
via a footnote. A short version of the title, not exceeding 35 characters, should be provided for a running 
head. An abstract of about 250 words should accompany all research articles on a separate page. 

Tables, one to a page, should be double-spaced throughout and be assigned consecutive Arabic numer- 
als. Collect all figure legends on a separate page. Each illustration should be centered on a single page 
and be no smaller than final size and no larger than twice final size. The name of the author(s) and figure 
number, assigned consecutively using Arabic numerals, should be pencilled on the back of each figure. 

Names for birds should follow the A.O.U. Checklist of North American Birds (7th ed., 1998) or another 
authoritative source for other regions. Subspecific identification should be cited only when pertinent to 
the material presented. Metric units should be used for all measurements. Use the 24-hour clock (e.g., 
0830 H and 2030 H) and “continental” dating (e.g., 1 January 1999). 

Refer to a recent issue of the journal for details in format. Explicit instructions and puhlication policy 
are outlined in “Information for contributors,”/. Raptor Res., Vol. 39(4), and are available from the editor. 
Submit manuscripts to incoming Editor-in-chief Cheryl Dykstra, Raptor Environmental, 7280 Susan Springs 
Drive, West Chester, OH 45069-3696 U.S.A.journalofraptorresearch@juno.com 

COVER: Cuban Black-Hawk {Buteogallus anthracinus gundlachii). Painting by Nils Navarro. 


Contents 

Taxonomic Status and Biology of the Cuban Black-Hawk, Buteogallus 

ANTHRACINUS GUNDLACHII (AVES: AcCIPITRIDAE) . James W. Wiley and Orlando H. Garrido 351 I 

Home Range and Habitat Use of Northern Spotted Owls on the Olympic 

Peninsula, Washington. Eric D. Forsman, TimmothyJ. Kaminski, Jeffery C. Lewis, 

Kevin J. Maurice, Stan G. Sovem, Cheron Ferland, and Elizabeth M. Glenn 365 

First-cycle Molts in North American Falconieormes. Peter Pyie 378 

Morphometric Analysis of Large Falco Species and Their Hybrids with 

Implications for Conservation. Chris p. Eastham and Mike k Nichoiis 386 

A Change in Foraging Success and Cooperative Hunting by a Breeding Pair 

of Peregrine Falcons and Their Fledglings. Dick Dekker and Robert Taylor 394 

Nesting Ecology and Behavior of Broad-winged Hawks in Moist Karst 

Forests of Puerto Rico. Derek W. Hengstenberg and Francisco J.VUella 404 

Raptor Abundance and Distribution in the Llanos Wetlands of Venezuela. 

Wendy J. Jensen, Mark S. Gregory, Guy A. Baldassarre, Francisco J. Vilella, and Keith L. Bildstein .... 417 

A Comparison of Breeding Season Food Habits of Burrowing Owls Nesting 
IN Agricultural and Nonagricultural Habitat in Idaho. Coiieen e. Moulton, 

Ryan S. Brady, and James R. Belthoff 429 

Red-tailed Hawk Dietary Overlap with Northern Goshawks on the Kaibab 

Plateau, Arizona. Angela E. Gatto, Teryl G. Grubb, and Carol L. Chambers 439 

Bat Predation by Long-eared Owls in Mediterranean and Temperate Regions 

OF Southern Europe. Ana Maria Garcia, Francisco Cervera, and Alejandro Rodriguez 445 

Short Communications 

Differential Effectiveness of Playbacks for Little Owls (Athene noctua) Surveys Before and 
After Sunset. Joan Navarro, Eduardo Mmguez, David Garcia, Carlos Villacorta, Francisco 
Botella, Jose Antonio Sanchez-Zapata, Martina Carrete, and Andres Gimenez 454 

King Vultures (Sarcoramphus papa) Forage in Moriche and Cucurit Palm Stands. Marsha A. Schlee 458 

Family Break Up, Departure, and Autumn Migration in Europe of a Family of Greater Spotted 
Eagles (Aquila clanga) as Reported by Satellite Telemetry. Bernd-U. Meyburg, Christiane 
Meyburg, Tadeusz Mizera, Grzegorz Maciorowski, and Jan Kowalski 462 

Seasonal Patterns of Common Buzzard (Buteo buteo) Relative Abundance and Behavior in ^ 

POLLINO National Park, Italy. Massimo Pandolfi, Alessandro Tanfema, and Giorgia Gaibani ...... 466 

New Nesting Record and Observations of Breeding Peregrine Falcons in Baja C al ifornia Sur, 

Mexico. Aradit Castellanos, Cerafina Arglielles, Federico Salinas-Zavala, and Alfredo Ortega-Rubio 472 

Letters 

A Previously Undescribed Vocalization of the Northern Pygmy-Owl. Graham G. Frye 476 

Book Review. Edited byjoelle Gehring 478 

Information for Contributors 480 

Index to Volume 39 484 


The Raptor Research Foundation, Inc. gratefully acknowledges funds and logistical support 
provided by Arkansas State University to assist in the publication of the journal. 


THE JOURNAL OF RAPTOR RESEARCH 

A QUARTERLY PUBLICATION OF THE RAPTOR RESEARCH FOUNDATION, INC. 


VoL. 39 December 2005 No. 4 


J. Raptor Res. 39(4) :351— 364 
© 2005 The Raptor Research Foundation, Inc. 


TAXONOMIC STATUS AND BIOLOGY OF THE CUBAN BLACK- 
HAWK, BUTEOGALLUS ANTHRACINUS GUNDLACHII 

(AVES: ACCIPITRIDAE) 

James W. Wiley^ 

USGS Maryland Cooperative Fish and Wildlife Research Unit, University of Maryland Eastern Shore, 

1120 Trigg Hall, Princess Anne, MD 21853 US. A. 


Orlando H. Garrido 

1706 Calle 60 entre 17 y 19, Marianao 13 (Play a), La Habana, Cuba 

Abstract. — ^We reevaluate the taxonomic status of the Cuban population of the Common Black-Hawk 
{Buteogallus anthracinus) based on our examination of additional specimens, nests, eggs, and voice data. 
Buteogallus a. gundlachii is smaller than mainland populations of anthracinus and differs from mainland 
birds in plumage coloration and pattern. The common (alarm) call of gundlachii is a series of three or 
four notes, differing from that of mainland anthracinus, whose call consists of 9-24 notes. In the Isla de 
Pinos, Cuba, we observed gundlachii ediiing two species of land crabs (71.4%), centipedes (7.1%), lizards 
(10.7%), mammals (7.1%), and a bird (3.6%). We consider Buteogallus gundlachii Cabanis 1854 (1855), 
the Cuban Black-Hawk, to be a full species, endemic to Cuba, Isla de Pinos, and many of the cays of 
the Cuban Archipelago. 

Key Words: Common Black-Hawk; Buteogallus anthracinus; Cuban Black-Hawk; Buteogallus gundlachii; 

Buteogallus subtilis; ecology; taxonomy. 


ESTADO TAXONOMICO Y BIOLOGIA DE BUTEOGALLUS ANTHRACINUS GUNDLACHII (AVES: 
ACCIPITRIDAE) 

Resumen. — En este estudio re-evaluamos el estatus taxonomico de la poblacion cubana de Buteogallus 
anthracinus (subespecie gundlachii) con base en examenes de especimenes adicionales, nidos, huevos y 
datos de la voz. Los individuos de B. a. gundlachii son mas pequehos que los individuos de las poblaciones 
continentales de B. anthracinus, y difieren de las aves del continente en la coloracion y patron del 
plumaje. El llamado comun de alarma de gundlachii es una serie de tres o cuatro notas, mientras que 
el llamado de anthracinus en el continente consiste de entre 9 y 24 notas. En la Isla de Pinos, Cuba, 
observamos a gundlachii alimentandose de dos especies de cangrejos terrestres (71.4%), ciempies 
(7.1%), lagartijas (10.7%), mamiferos (7.1%) y un ave (3.6%). Consideramos Buteogallus gundlachii 
Cabanis, 1854 (1855) debe ser tratado como una especie distinta, endemica de Cuba, la Isla de Pinos 
y muchos de los cayos del archipielago cubano. 

[Traduccion del autores] 


The New World genus Buteogallus Lesson, 1830 
includes five species, mostly restricted to tropical 
areas, including Great Black-Hawk {Buteogallus uru- 


^ Email address: jwwiley@mail.umes.edu 


bitinga) of the lowlands of Mexico to northern Ar- 
gentina, Savanna Hawk {B. meriodionalis) inhabit- 
ing savannas and marshes of western Panama to 
northern Argentina, Rufous Crab-Hawk (B. aequi- 
noctialis) occurring in mangroves of northeastern 


351 


352 


Wiley AND Garrido 


VoL. 39, No. 4 


Venezuela to eastern Brazil (Parana), and Man- 
grove Black-Hawk {B. subtilis), which is restricted 
to the Pacific coasts and rivers of El Salvador south 
to northwestern Peru. The Common Black-Hawk 
(Buteogallus anthracinus, Deppe 1830) ranges from 
southwestern United States, south to extreme 
northern South America (coastal Venezuela to 
northeastern Guiana) , Colombia, to northern 
Peru, including Trinidad, and some of the West 
Indies (Bond 1950, American Ornithologists’ 
Union 1998). One of the West Indian populations 
{B. a. cancrivorus Clark 1905b) is restricted to St. 
Vincent, St. Lucia, Union Island (Grenadines), and 
Grenada (accidental and doubtful in last two is- 
lands; no specimens taken; Clark 1905a, 1905b, 
1905c, Bond 1950, Evans 1990) in the Lesser An- 
tilles, whereas the only other Antillean population 
{B. a. gundlachii Cabanis 1854 [1855]) occurs in 
Cuba and its satellites. The taxonomic status of the 
Cuban population has been controversial, with 
some considering the form as a full species, Buteo- 
gallus gundlachii (as originally described by Cabanis 
[1854, actually 1855]) instead of Buteogallus anth- 
racinus gundlachii (American Ornithologists’ Union 
1998) . Among those authorities who have consid- 
ered the Cuban form gundlachii conspecific with 
the continental species {anthracinus) are Sharpe 
(1874, 1899), Cory (1887, 1892), Bangs and Zap- 
pey (1905), Bond (1956a, 1956b), Amadon (1961), 
Brown and Amadon (1968), Mayr and Short 
(1970), Stresemann and Amadon in Mayr and Cot- 
trell (1979), Palmer (1988), Sibley and Monroe 
(1990), Ferguson-Lees and Christie (2001), Dick- 
inson (2003), and others. Conversely, other au- 
thors have considered gundlachii different from Bu- 
teogallus anthracinus at the species level: Cabanis 
(1855), Gundlach (1854, 1865-1 866a, 1865-1 866b, 
1871, 1876), Ridgway (1876), Gurney (1876, 1934), 
Bangs (1905), Swann (1921-1922, 1930), Peters 
(1931), Bond (1936), Hellmayr and Conover 
(1949), Friedmann (1950), Monroe (1963, 1968), 
Wetmore (1965), and others. Some of these au- 
thors subsequently changed their opinions on the 
Cuban form’s status, later considering gundlachii 
conspecific with anthracinus (e.g., Gundlach 1893, 
Bond 1950, 1956a, 1956b). With rare exception, 
however, previous evaluations did not consider the 
important characteristics of breeding biology and 
voice, mainly because of the limited knowledge of 
the Cuban form resulting from the difficulty in 
reaching its breeding habitats. The lack of natural 
history information is not unique to gundlachii, but 


is also true for other forms of the genus Buteogallus, 
e.g., anthracinus and subtilis, which are currently 
recognized as different species (Aldrich and Bole 
1937, Amadon 1982, Mayr and Cottrell 1979, Stiles 
and Skutch 1989, Sibley and Monroe 1990, Amer- 
ican Ornithologists’ Union 1998, Ridgely and 
Greenfield 2001), but with reservation by some au- 
thors (Stiles and Skutch 1989, American Ornithol- 
ogists’ Union 1998, Ridgely and Greenfield 2001). 

Here, we reevaluate the taxonomic status of the 
Cuban population of Buteogallus anthracinus gund- 
lachii, based on our examination of more speci- 
mens, nests, eggs, and behavioral data, especially 
vocalizations, than were considered by previous 
workers. All published work on the Cuban form 
has been based on information from the few spec- 
imens collected before 1960, all of which are de- 
posited in foreign institutions. In this study, we in- 
clude specimens in North American and Cuban 
collections, including those collected after 1960, 
and not evaluated previously. 

Our main comparison in this assessment is with 
anthracinus, the taxon most often linked to gund- 
lachii. In these comparisons, we refer to Cuban 
populations as gundlachii and other forms as anth- 
racinus. It is not the purpose of this contribution 
to speculate on the taxonomic status of Buteogallus 
subtilis (including the three subspecies), although 
we make some comparisons between subtilis and 
gundlachii. 

Study Area 

Many of the observations reported here were made 
during our 30 yr of field experience throughout Cuba. 
We made more intensive observations of nesting black- 
hawks from 1996 to 1998 at the Los Indios Ecological 
Reserve, Isla de Pinos (now Isla de la Juventud). M^or 
vegetational communities at Los Indios include: (1) man- 
grove forest formation, characterized by black mangrove 
(Avicennia germinans) and red mangrove {Rhizophora man- 
gle); (2) semi-deciduous gallery forests, with prominent 
Cuban royal palm (Roystonea regia), beach hibiscus {Ht- 
hiscus tiliaceus), and pond apple {Annona glabra); (3) the 
open forest (savanna) formation of an open pine {Pinus 
caribaea and P. tropicalis) and Cuban bottle palm ( Colpoth- 
rinax rmightii) , with silver saw palm (Acoelorraphe wrightn) 
and a sparse undergrowth; and (4) the pine-barren for- 
mation, with pines and palms, and an undergrowth pre- 
dominantly of Pachyanthus cubensis, P. poiretii, Kalmiella ag- 
gregata, Miconia delicatula, Polygala uncinata, Lyonia 
myrtilloides, and Pinguicula filifolia (Jennings 1917, Alain 
1946). Black-hawk observations were made mainly in the 
mangrove and gallery forests. Additional intensive obser- 
vations in red and black mangrove habitats were made 
in Cienaga de Zapata in December 1999. An elevated 
road bed, lined with Casuarina equisetifolia and scrub veg- 


December 2005 


Cuban Black-Hawk Status and Biology 


353 


elation, bisects the mangrove forest where we made our 
observations near Playa Larga. 

Methods 

We examined specimens of Buteogallus a. anthracinus 
(N = 37), B. a. gundlachii (12), B. a. cancrivorus (4), B. 
subtilis (25), B. aequinoctialis (3), and B. urubitinga (24) 
deposited in the Field Museum of Natural History (Chi- 
cago), Museum of Comparative Zoology (Harvard Uni- 
versity), American Museum of Natural History, United 
States National Museum of Natural History, Academy of 
Natural Sciences of Philadelphia, Louisiana State Univer- 
sity Museum of Natural History, Museo Nacional de His- 
toria Natural de Cuba (La Habana), and Instituto de Eco- 
logia y Sistematica (Cuba) (Table 1). Conventional 
measurements of wing chord (flattened against the rul- 
er), tail, tarsus, and exposed culmen were taken to the 
nearest 0.1 mm with calipers. Egg masses were measured 
to the nearest gram using spring scales. We present sum- 
mary descriptive statistics (mean, SD, and range) for the 
specimens. We plotted body measurements to assess the 
pattern of spatial segregation among populations and 
forms. The hypothesis of separation derived from the 
plots of body measurements was tested using discriminate 
function analysis (DFA; Kleinbaum and Kupper 1978). 
SPSS (1999) for Windows was used to run DFA. 

Results 

Morphometries and Plumage. Adult morphology. 
Our examinations of the two taxa of B. anthracinus 
(anthracinus and gundlachii) revealed differences in 
size and coloration. We found sexual size dimor- 
phism in three of the measurements taken of spec- 
imens of mainland anthracinus (Table 1). There- 
fore, size comparisons between anthracinus and 
gundlachii were made within sex; i.e., male anthra- 
cinus with male gundlachii and female anthracinus 
with female gundlachii. Tarsal length was not dif- 
ferent in either population, so for comparing anth- 
racinus with gundlachii tarsi we combined male and 
female measurements for that morphometric pa- 
rameter. Only measurements of wing and exposed 
culmen for gundlachii revealed sexual size dimor- 
phism (P < 0.01; Table 1), although the small sam- 
ple size of females (N — 5) precluded a reliable 
analysis. Measurements of gundlachii yielded a 
mean Dimorphic Index (Storer 1966) of 6.9, com- 
pared with a mean index of 5.6 for anthracinus (Ta- 
ble 1). 

Birds from Cuba (gundlachii) are substantially 
smaller than mainland (anthracinus) birds in some 
conventional measurements, including wing chord 
in both sexes and tail length in males (Table 2). 
Also, tarsal lengths (combined male and female 
measurements) were significantly different be- 
tween the two forms (P — 0.001). A stepwise selec- 


tion procedure within DFA revealed wing chord, 
tail length, and exposed culmen were the most im- 
portant of the size variables measured. Plots con- 
trasting these variables within sex showed anthra- 
cinus and gundlachii tended to occupy generally 
distinct regions of the morphological space (Fig. 
1 ). 

To further examine size differences between the 
two populations, we used linear discriminant anal- 
ysis to classify specimens into two groups (“race”), 
mainland anthracinus and Cuban gundlachii, using 
lengths of wing chord, tail, culmen, and tarsus as 
predictors. For male anthracinus, the analysis pro- 
duced a true group classification proportion of 
0.938 (15 of 16 correctly classified) and 0.857 (6 
of 7 correctly classified) for gundlachii males, for 
an overall proportion correct of 0.913 (21 of 23) 
(Wilks’s lambda = 0.375; = 17.646; df = 4, P< 

0.001). For females, the analysis produced a true 
group classification proportion of 0.857 (18 of 21) 
for anthracinus and 0.800 (4 of 5) for gundlachii 
individuals, for an overall proportion correct of 
0.846 (22 of 26) (Wilks’s lambda = 0.4.95; = 

14.781.1; df = 4, P< 0.005). 

The four adult female St. Vincent (B. a. canen- 
vorus) specimens we examined were somewhat 
larger in wing chord (x = 389 ± 7.63, range = 
385-401; t = -4.99, P = 0.002, df = 6) than gund- 
lachii females, whereas we found no difference be- 
tween the two island forms in tail (213.3 ± 12.4; 
range = 200-230; t - -1.83, P > 0.05, df = 6), 
culmen (27.3 ± 0.8; range = 26.8-28.4; t = 0.60, 
P > 0.05, df = 5), or tarsus (85.5 ± 6.4; range = 
81.0-94.9; t = -1.63, P> 0.05, df = 4) length. We 
found no differences (P > 0.05) in measurements 
between anthracinus and cancrivorus. 

In general coloration, gundlachii differs from B. 
anthracinus and B. subtilis in being chocolate- 
brown, not slate blackish or even black as in the 
latter two forms. However, some specimens of anth- 
racinus, especially of the race cancrivorus, have a 
tendency to be less blackish, almost dark brown. 

The underparts feathers of gundlachii have a 
light (brownish-gray) edge, more conspicuous to- 
ward the abdominal region and more broadly 
edged on the alula coverts than in anthracinus, with 
the edging on the terminal alula coverts becoming 
white bands. The margins of the flank and thigh 
feathers are heavily marked, forming a series of 
bands, although these bands tend to disappear in 
older birds. The shoulder feathers are boldly 
barred in white, contrasting with the chocolate- 


354 


Wiley AND Garrido 


VoL. 39, No. 4 


Table 1. Sexual size dimorphism in four body measurements from specimens of Buteogallus anthracinus (mainland 
Buteogallus a. anthracinus -aind Cuban B. a. gundlachii) , B. subtilis, B. aequinoctialis, and B. urubitinga, expressed as mean, 
standard deviation, range, and sample size (in parentheses). Statistical analyses are between-sex comparisons (two- 
sample t-test; equal variances not assumed) . 


Species 

Structure 

Sex 

Male 

Female 

t 

df 

P 

Signif- 

icance® 

D.l.^ 

Buteogallus anthracinus 








Wing 

371.69 ± 11.95 (16) 

385.19 ± 11.21 (21) 

-3.50 

31 

0.001 

* 

3.6 


341-393 

360-421 






Tail 

195.50 ± 7.40 (16) 

213.81 ± 10.61 (21) 

-6.18 

34 

0.0001 


8.9 


183-210 

190-230 






Exposed culmen 

26.27 ± 0.82 (16) 

27.40 ± 1.25 (20) 

3.24 

32 

0.003 


4.2 


25.1-28.1 

23.6-30.3 






Tarsus 

85.94 ± 2.65 (16) 

85.42 ± 4.00 (21) 

0.48 

34 

0.636 

ns 

-0.6 


80-90.0 

80.3-92.7 






Mean D.I. 







4.0 

Buteogallus gundlachii 








Wing 

342.71 ± 12.16 (7) 

363.00 ± 8.43 (5) 

-3.41 

9 

0.008 

* 

5.8 


323-370 

350-372 






Tail 

179.29 ± 9.12 (7) 

191.60 ± 22.16 (5) 

-1.15 

4 

0.313 

ns 

6.6 


167-197 

182-233 






Exposed culmen 

25.32 ± 0.69 (7) 

27.54 ± 0.61 (5) 

-5.84 

9 

0.0001 


8.3 


24.5-28.5 

26.7-28.1 






Tarsus 

81.33 ± 3.57 (6) 

79.67 ± 3.56 (5) 

0.77 

8 

0.464 

ns 

-2.1 


75.4-87 

79.0-87.7 






Mean D.I. 







4.7 

Buteogallus subtilis 








Wing 

348.0 ± 13.68 (12) 

352.31 ± 13.43 (13) 

-0.79 

22 

0.436 

ns 

1.2 


330-370 

328-373 






Tail 

189.92 ± 14.58 (12) 

191.69 ± 8.65 (13) 

-0.37 

17 

0.719 

ns 

0.9 


168-220 

180-205 






Exposed culmen 

25.46 ± 2.06 (12) 

26.56 ± 1.56 (12) 

-1.47 

20 

0.157 

ns 

4.2 


19.7-27.8 

23.1-28.8 






Tarsus 

79.78 ± 3.09 (11) 

79.55 ± 2.76 (13) 

0.19 

19 

0.848 

ns 

-0.3 


73.3-84.1 

75.0-84.0 






Mean D.I. 







1.5 

Buteogallus aequinoctialis 







Wing 

315.5 ± 0.71 (2) 

322 (1) 







315-316 







Tail 

155.0 ± 2.83 (2) 

155 (1) 







153-157 







Exposed culmen 

23.55 ± 0.92 (2) 

16.8 (1) 







22.9-24.2 







Tarsus 

74.5 ± 3.54 (2) 

72.8 (1) 







72-77 







Buteogallus urubitinga 








Wing 

384.94 ± 16.68 (16) 

389.63 ± 18.36 (8) 

-0.61 

12 

0.555 

ns 

0.1 


362-412 

365-415 






Tail 

225.13 ± 13.50 (16) 

234.63 ± 17.54 (8) 

-1.35 

11 

0.206 

ns 

4.1 


190-250 

210-260 






Exposed culmen 

29.72 ± 1.05 (16) 

30.78 ± 2.20 (8) 

-1.84 

8 

0.103 

ns 

3.5 


26.7-30.9 

27.2-34.2 






Tarsus 

112 ± 8.25 (16) 

113.16 ± 7.76 (8) 

-0.11 

14 

0.910 

ns 

1.0 


85.9-118.9 

98.8-123.0 






Mean D.I. 







2.2 


® Significance, * = P < 0.05, ns = not significant. 
D.I. = Dimorphic Index (Storer 1966). 


December 2005 


Cuban Black-Hawk Status and Biology 


355 


Table 2. Mean, standard deviation, and sample size (parentheses) for wing chord, tail, culmen, and tarsus length 
for mainland {Buteogallus a. anthracinus) and Cuban {Buteogallus a. gundlachii) populations of the Common Black- 
Hawk. Statistical analyses are within-sex comparisons (two-sample <-test; equal variances not assumed) between main- 
land and Cuban specimens, except for tarsus, for which we found no sexual size dimorphism. 


Structure 

Sex 

Taxon 

B. A. ANTHRACINUS B. A. GUNDLACHII 

t 

df 

P 

Signieicance^* 

Wing 

M 

371.69 ± 11.95 (16) 

342.71 ± 12.16 (7) 

5.28 

11 

<0.001 



F 

385.19 ± 11.21 (21) 

363.00 ± 8.43 (5) 

4.94 

7 

0.002 

SH 

Tail 

M 

195.50 ± 7.40 (16) 

179.29 ± 9.12 (7) 

4.14 

9 

0.003 

:)4 


F 

213.81 ± 10.61 (21) 

191.60 ± 22.16 (5) 

2.14 

4 

0.099 

ns 

Exposed 

M 

26.27 ± 0.82 (16) 

25.32 ± 0.69 (7) 

2.86 

13 

0.013 


culmen 

F 

27.40 ± 1.25 (20) 

27.54 ± 0.61 (5) 

-0.36 

13 

0.728 

ns 

Tarsus 

M and F’’ 

85.64 ± 3.45 (37) 

80.57 ± 3.49 (11) 

4.24 

16 

0.001 



® Significance, * = P < 0.05, ns = not significant. 

^ Male and female tarsus data combined because specimens did not display sexual size dimorphism. 


brown ground color. Remiges are dark brown, with 
wing coverts edged in grayish-cinnamon, especially 
the secondaries. The undersides of primaries and 
some secondaries have an extensive white patch, 
which constitutes the most distinctive character of 
the Cuban form. In subtilis, and especially anthra- 
cinus, this patch is mottled with grayish-brown. The 
tertiaries of gundlachii are heavily mottled grayish. 
This mottling is similar to the coloration of the 
primaries and secondaries of anthracinus, which 
has only an inconspicuous whitish patch on the un- 
dersides of these feathers. On the other hand, 
some specimens of subtilis display more white in 
this region than does anthracinus, but do not ap- 
proach the amount shown in gundlachii. 

The upperparts in gundlachii are also brown, 
with brownish-gray or with a trace of cinnamon on 
the feather margins. The head and pileum are uni- 
formly chocolate brown. The rectrices are darker 
brown, almost blackish, with a broad white band 
of variable width (averaging 40 mm) in the middle 
of the tail. The tip of the tail is edged in white (as 
wide as 13 mm), which is a purer white than in 
anthracinus and subtilis. The feet and cere are yel- 
low, the claws are black, and the iris is dark brown. 
The bill is blackish at the tip, becoming more yel- 
lowish toward the base on maxilla and mandible. 

Immature morphology. Immature gundlachii individ- 
uals are not chocolate brown ventrally, but rather 
whitish, and heavily mottled with brown, having 
some feathers with considerable beige suffusion. 
Many feathers are mottled with medallion-like 
marks, whereas others are marked with elongated 
blotches, and some with streak-like dashes; these 


marks are seldom present in fully-feathered im- 
mature birds. The sides of the face and throat are 
whitish, speckled with hrown. The pileum, nape, 
and neck are heavily mottled or spotted with 
brown on a light (white or beige) background. 
Flanks and thighs also display considerable varia- 
tion, with younger birds showing a lighter (whitish 
to brownish-beige) background, whereas older 
birds display more mottling or barring. The thighs 
are distinctly barred with light and dark bands in 
subtilis and anthracinus, whereas gundlachii has mot- 
tled or very lightly barred thighs. 

The white patch of the underside of primaries is 
even more expanded and conspicuous in subadult 
than in adult gundlachii. Also, the subadult’s tail is 
distinct from that of the adult’s tail. When still not 
in full adult plumage, the subadult’s tail shows 
remnants of several (as many as nine) thin, brown- 
ish bands, instead of displaying a single broad 
white band in the middle of the tail as in the adult. 
Some bands are complete, whereas others are 
somewhat broken. In Cuban birds, these bands are 
straight and parallel, whereas in the other forms 
they are oblique (chevron-like), as well as being 
much wider than in gundlachii. The bands become 
less delimited toward the tip; compared with the 
adult, the white tip of the subadult’s tail is less de- 
marcated, more grayish than white, and becomes 
browner from the tip toward the base. 

Natural History. Habitat. Although we occasion- 
ally observed black-hawks within the white sand 
palm savanna of Los Indios, Isla de Pinos, nearly 
all observations were made in the coastal zone, pri- 
marily in mangrove forests or at the edges of that 


356 


Wiley AND Garrido 


VoL. 39, No. 4 


Male 



Female 



420 
410 
400 
O) 390 
5 380 

370 
360 
350 

23.5 24.5 25.5 26.5 27.5 28.5 29.5 30.5 23.5 24.5 25.5 26.5 27.5 28.5 29.5 30.5 

Culmen Culmen 


230 
220 
210 

_ 200 
<0 

^ 190 

180 
170 
160 

75 80 85 90 

Tarsus 

Figure 1. Plots contrasting body measurements of specimens of mainland Buteogallus anthracinus anthradnus (solid 
dots; = 16 males, 21 females) and Cuban B. a. gundlachii (open circles; N = 7 males, 5 females). 






habitat. Hawks hunted in the sparsely-vegetated 
mangrove pannes and flooded openings, where 
they foraged by perching in young or dead man- 
groves. We also saw black-hawks foraging or roost- 
ing in beach and coastal habitats, frequently perch- 
ing in windbreaks of Casuarina equisetifolia at the 
edge of mangroves and dirt roads. 

Nidification. We examined eight nests at Los In- 
dios within the period of 14-27 May 1996-98. All 


contained eggs, except the nest examined on 27 
May 1996, which had one chick. During our obser- 
vations at Isla de Pinos, which were well into the 
breeding season, we observed no aerial courtship, 
although individual gundlachii regularly soared si- 
lendy for short periods above their nesting areas. 

Of the eight gundlachii nests we examined at Los 
Indies, half were placed in black-mangroves and 
half in red-mangroves (Table 3) . Each of the nests 


Table 3. Nest and contents data for eight Buteogallus anthracinus gundlachii nests at Los Indios, Pinos, Cuba, 1996-98. 


December 2005 


Cuban Black-Hawk Status and Biology 


357 


00 


00 

03 

O) 


CM 


CM 


J> 

03 

O) 


pa 

S 

H 

z 


oo 


to 

03 

03 


CM 


I 

I- 


00 m 
00 


CM 




CO 

CM 

X 

UO 


to 


CO 

g CM 

S 

X 

C.O) th fti 

pS o 03 I-H 30 
^ ^ CM 3T3 


O TfH 

to to 


ts 

* 

i 


ts 

■r-i 

c 

' C»i 


ts 

• esi 

S 

I 

s> 


CM CM 
GO CM 

X X 


30 lO o 

O 03 ^ 


^ CM CO 

to to to to 


CM 30 30 


30 


O 30 CM 
00 i> ^ 


w 


CO CN| 

X X 
00 

30 to 03 
30 30 30 


30 03 I-H 30 
03 03 I—* 


to 03 
GO CM 

X X 
w ^ ^ 

to Th CO 03 
CM 30 30 to 30 


i 


Tf oo 03 Tft 
03 00 r-H 


I 

t- 

■S3 30 


CO 00 00 30 
03 00 rH 


U 


rH 03 

CM 1 — i 
Tft Tft 

X X 

^O CO 

30 30 CT3 O 
CM 30 30 30 to 




s 

s 

•r* 


O 30 
1> 




CM 

'tfi 

X 

CO 

to 

30 


to 




? 


13 

H 

Z 

W 

•G S 

U 

B 

ij 

ji 

(X 

Z 

lU 

<u 


13 

0 

Qh 

-i-i 

<u 



0, 

^ Xi 



s 

qj be 

§ 
• ^ 

Ph 


0 

u 'G 

13 

0 

u 

b -c 



X! 





■t-i 


[fi !/3 

Vi 

«3 

V} 


13 13 


13 

13 


o 


CM 

6 

z 


o 

Z 


CM 

6 

Z 


z z z z z 


c 

o 

4-> 

a 

o 

O 


CO 

N 


bO 

be 

W 


VI 

m 

d 


bc 

be 

W 


« 


I 

f 




ts 

s 


be 

ts 

•«»» 

s 

s 

*<«* 


U 

U 


U 


G 

§ 


bio 

V 


.o 


was in the subcanopy, shaded by foliage and was 
constructed completely of Avicennia and Rhizophora 
twigs. The nests showed a large range of sizes, 
probably the result of additions made in successive 
years. Two of the nests we monitored from 1996 
through 1998 were reused by black-hawks, and in- 
creased in size with the addition of more nest ma- 
terials in subsequent years. All nests examined at 
Los Indios contained fresh or older lining materi- 
als, consisting of green leaves and sprigs of Avicen- 
nia and Rhizophora, and some debris. Both adults 
were observed bringing green lining material to 
nests. 

Nests had notably deep bowls (Table 3) and 
when adults were on nests incubating or brooding, 
they remained low in the bowl and were difficult 
to detect. During our inspections of nests at Los 
Indios, adults at three nests regularly perched plac- 
idly within 2 m of us while we measured eggs and 
chicks. Adults at a fourth nest were somewhat more 
aggressive, but the pair only flew low above our 
heads, occasionally calling, and vocalized from a 
nearby perch while we measured eggs. 

We measured 11 eggs at Los Indios (Table 3). 
Three eggs collected by O. H. Garrido in Cayo 
Candles (Archipielago de los Canarreos; deposited 
at Instituto de Ecologia y Sistematica) measured 
55.16 X 44.1 mm, 55.8 X 42.6 mm, and 57.08 X 
42.34 mm. The 14 gundlachii eggs we measured av- 
eraged 55.87 ± 0.69 (range = 54.7-57.08) X 42.71 
± 0.62 (range = 41.9—44.1) mm. Eggs of gundla- 
chii are typically short sub-elliptical to elliptical, 
with a finely granulated texture. Eggs have a dull 
grayish-white ground color, sometimes with a 
greenish or bluish cast early in incubation, and are 
marked with spots and blotches of dark or reddish- 
brown, particularly at the larger end. Clutch sizes 
at Los Indios averaged 1.57 ± 0.53 {N — 8; range 
= 1-2) eggs (Table 3). The egg of gundlachii is 
usually more colored (bluish to greenish suffu- 
sion) than those of anthracinus or subtilis, which are 
typically grayish or whitish (Bent 1937, Wetmore 
1965, O. Garrido pers. obs.). 

Diet and foraging behavior. Cuban birds were 
found to feed on a variety of prey (Table 4) . No- 
table was the lack of fish prey, although fishes were 
available in tidal channels in the study area. How- 
ever, twice, hawks were observed wading in shallow 
tidal channels and making foot thrusts at probable 
fish prey. During our observation periods (May- 
June) at Los Indios, land crab populations were 
particularly high, and crabs were active and con- 


358 


Wiley AND Garrido 


VoL. 39, No. 4 


Table 4. Prey of Buteogallus anthracinus gundlachii at Los Indios, Isla de Pinos, Cuba, 1996—1998, and Cienaga de 
Zapata, Cuba, 1999-2000. 


Prey 



Number (%) 


Observed Brought 
TO Nest 

Prey Remains 

Observed Captures 

Total (%) All 
Observations 

Invertebrates 





Crab 





Cardisoma guanhumi 

4 

12 

2 

18 (64.3) 

Ucides cordatus 

1 

1 


2 (7.1) 

Centipede sp. 

1 

1 


2 (7.1) 

Totals (invertebrates) 

6 (21.4) 

14 (50.0) 

2 (7.1) 

22 (78.6) 

Vertebrates 





Reptiles 





Lizards 





Anolis spp. 

1 

1 


2 (7.1) 

Ameiva auberi 

1 



1 (3.6) 

Totals (reptiles) 

2 (7.1) 

1 (3.6) 


3 (10.7) 

Birds 





Sora Porzana Carolina 


1 


1 (3.6) 

Totals (birds) 


1 (3.6) 


1 (3.6) 

Mammals 





Rattus rattus 


2 


2 (7.1) 

Totals (mammals) 


2 (11) 


2 (7.1) 

Total (vertebrates) 

2 (7.1) 

4 (14.3) 


6 (21.4) 

Total (all observations) 

8 (28.6) 

18 (64.3) 

2 (7.1) 

28 


spicuous in the early mornings and evenings, when 
most of our observations of prey delivery and cap- 
tures were made. In December 1999, we also ob- 
served gundlachii capturing several crabs {Cardiso- 
ma guanhumi) along the coast of Cienaga de 
Zapata, where the hawks hunted from a mixed 
mangrove- equisetifolia-codistal scrub zone. 

During our observations in the Los Indios man- 
grove habitat, gundlachii displayed passive still 
hunting from low (x = 1.3 ± 0.94; range = 0.2-3 
m; N = 54) mangrove tree perches or from the 
ground. Prey captures were made in a low-angle 
flight, snatching the item (all observations of 
crabs) and continuing to a nearby perch, or the 
hawk landed near the crab and stalked it on foot. 
Once the hawk grasped the crab, it controlled the 
claws and legs on either side of the prey with its 
feet, then removed the carapace with a quick tug 
at the head region using its bill. 

We found apparent caches of uneaten, though 
dismembered, land crabs near (range = 5-20 m) 
used gundlachii nests. However, we did not observe 


hawks returning to the caches to feed on the stock- 
piled crabs. 

Although B. a. anthracinus has been observed 
(O. Garrido pers. obs.) in Mexico hunting at the 
edge of a meadow in a fashion similar to that of 
the coursing behavior of the Northern Harrier 
{Circus cyaneus), gundlachii was not observed for- 
aging aerially in an active manner. 

Vocal behavior. The common call of gundlachii is 
a series of three or, uncommonly, four notes, with 
emphasis on the first two elements, suggesting its 
Cuban common name, BA-TIS-ta (Gundlach 1893, 
Garrido and Schwartz 1969, Garrido and Kirkcon- 
nell 2000; Fig. 2A). The call has a much shorter 
duration and fewer elements than in other popu- 
lations of Buteogallus anthracinus (Table 5). The 
common call of mainland anthracinus consists of 9- 
24 notes, with the middle to the final third of the 
notes accentuated (Fig. 2C-F, Table 5). Stiles and 
Skutch (1989) characterized the call of mainland 
anthracinus as ''klee klee klee KLEE KLEE klee kle kle 
keki ki." The comparable call of cancrivorus con&i&Xs, 


December 2005 


Cuban Black-Hawk Status and Biology 


359 




Figure 2. Sonographs of common (alarm) calls of Buteogallus species. A. Buteogallus anthradnus gundlachii, showing 
typical three element “ba-tis-ta” phrase, Cuba (G.B. Reynard), B, Buteogallus a. cancrivorus, St. Vincent (J. Roche, 
courtesy British Library Sound Archive) . C, Buteogallus a. anthradnus, Costa Rica (Cornell Library of Natural Sounds 
27216). D. Buteogallus a. anthradnus, Venezuela (R Schwartz). E. Buteogallus a. anthradnus, male, Arizona (courtesy J. 
Schnell). F. Buteogallus a. anthradnus, female, Arizona (courtesy J. Schnell). G. Buteogallus urubitinga, Venezuela (P. 
Schwartz) . H. Buteogallus aequinoctialis, Surinam (Paul Donahue, courtesy British Library Sound Archive) . 


Table 5. Characteristics {x ± SD, range in parentheses) of the common (advertisement) call of Buteogallus anthraanus anthraanus, B. a. cancnvorus, 
gundlachii, B. urubitinga, and B. aequinoctialis. 


Wiley AND Garrido 


VoL. 39, No. 4 


360 


53 


Q 

« 

s 

C/5 

H 

2 



o 

cO 

50 




w 

OI 

i> 

CM 

rH 

CM 



£ 

Ui 

S 

w 

h-i 

1 

1-H 

1 

50 

1 

00 

4i 

1 

05 

1 

1 

CM 












cO 

05 

50 

05 

M' 

00 

00 



00 

m 

CM 

CO 

00 

o 




I> 

GO 

I> 

CM 

o 


CO 




'Ch 

CO 


iD 

00 




xi 


1 

05 

Mi 

4 4 

1 

rH 



00 

rfH 


05 

o 

05 





CO 

CM 

05 

oo 

M' 

o 



oo 

OO 

CM 

CM 


CM 



O 

^ 





' 

' 


Hi 

X 

50 

rH 

OO 

50 

05 

CM 

CO 



OO 

GO 


d 

50 

d 




CO 

00 

^H 

05 

o 

M^ 

CO 



'Cfl 

GO 

50 

00 

oo 

CM 

rH 



+1 

+1 

+ 1 

+1 

+1 

+ 1 

+1 

N 












CM 


CM 

50 

CO 

o 




CD 

50 

CO 

50 





i> 

00 

o 

CO 

iD 

1-H 

rH 

G 


CO 

CO 

oo 

CO 


GO 


z 









w 









p 









a 




05 



V, 



CO 

GM 

1-H 

rH 

rH 

CM 


s 


Ol 

iC 

05 

1-H 

o 

CO 

CM 

50 



00 

rH 

rH 

CO 


00 




t 

1 






1 

00 

05 

1 

1 

1 

CN 



IM 

lO 

50 

oo 

o 

OO 

05 

05 



oo 

o 

OO 

00 


i'' 


> 

CO 

1-H 

rH 

50 

00 

CO 


0 





^H 

rH 

CM 



05 

05 

' ' 

' ' 





CO 

d 

00 

o 

oO 

00 

CO 



cb 

o 

05 

d 

d 

cb 

CM 



05 

CM 


CM 

05 

CO 

rH 



+ 1 

+l 

+1 

+ 1 

+ 1 

+ 1 

+ 1 



05 


CO 

M^ 



CM 



CO 

1-H 

CM 

05 

o 

CO 

50 



1> 

CO 

t£5 

50 

50 

j> 

cO 



1-H 

rH 

rH 

1-H 

rH 

1-H 

CM 














,^-v 

M-. 





lO 

CM 

CO 

GO 




z 

w 


rH 

1 

f-H 

1 


00 


ch 

1 

05 

rH 

rH 

1 

05 

CM 

CM 


1 

M^ 


V ^ 









o 


o 

05 

rH 

CO 

u 

o' 

CO 

50 

50 

50 

o 

d 

d 

+1 

+1 

+ 1 

+1 

+ 1 

+ 1 

+ 1 

Z 

1-H 

50 

CM 

50 

50 

o 

CO 


CO 

CM 

CO 

CM 

d 

1-H 

I> 



rH 

rH 

rH 

CM 

M^ 








CO 

s. 



05 

00 

oo 

50 

H 

q 



oo 

cb 

od 


1 

rH 

z 

0 

1 

00 

4^ 

1 

1 

05 

1 

o 

xO 

CM 

1 

05 

c 

d> 

CM 

CM 

1-H 

50 

rH 

d 

d 

'• — ^ 

' 


^ 



^ ^ 


^H 


00 

o 


CD 

rH 


o 

d 

rH 

rH 

rH 

rH 

d 

Q 

+1 

+1 

+ 1 

+1 

+1 

+1 

+ 1 


o> 

oo 

50 

CO 

00 

00 

o 


C5 

CO 


CM 

I> 

CM 

i-H 


J> 


50 

CM 

GO 

CO 

lO 





^ ^ 














— .. 

u 

r^ 






cd 

(2 

d 






0 

0 

u 






N 

C/D 

fi 






H 

0 







< 

u 

> 








'' 




0 

* Hi 
‘ Hi 

<< 

S 

s 

<<: 

S 

S 

‘G 

“3 

s 

fi 

H 






s 

b 

53 



■53 

e 

g 

1 

s 

s 

•Hi 

o 



s 

3 

S 

53 

?3 

53 

2 


■ri 

S 


<3 

53 

53 

53 

e 


53 


=q 

cq 

cq 

«sq 

oq 

=q 

oq 


cl 

o 


c 

u 

B 

t3 

a 

d 

ij-i 


of a large number of elements (22-37), with em- 
phasis on several middle elements (Fig. 2B, Table 
5). Similarly, the common call of B. subtilis is sub- 
stantially different from that of gundlachii, consist- 
ing of several, rapidly repeated elements, described 
by Ferguson-Lees and Christie (2001) as a series of 
shrill whistles, indistinguishable from anthracinus. 
The call of Buteogallus urubitinga consists of a single 
note, drawn out in a high shrill keeeeeeeeh" (Fer- 
guson-Lees and Christie 2001; Fig. 2G), whereas 
that of Buteo aequinoctialis is a distinct series of whis- 
tle-like notes (Ferguson-Lees and Christie 2001; 
Fig. 2H, Table 5). 

Discussion 

As is normal among most birds of prey, female 
gundlachii are somewhat larger than males, with 
culmen and wing length significantly different be- 
tween genders. Sexual size dimorphism was less ev- 
ident in anthracinus {x Dimorphic Index = 4.0) 
than gundlachii, where we found a mean Dimor- 
phic Index of 4.7 with males significantly larger 
than females in wing and culmen length (Table 1). 
Snyder and Wiley (1976) reported a lower index 
(2.7) of sexual size dimorphism for B. anthracinus. 

Whereas measurements of selected body parts 
did not show complete distinction between anth- 
racinus and gundlachii (Table 2, Fig. 1), Cuban 
birds were consistently smaller or at the small end 
of the range for anthracinus measurements. In con- 
trast to our measurements. Bangs (1905) partly 
based his determination of separating gundlachii 
from anthracinus on the former being slightly larg- 
er than the latter, and in having a decidedly heavi- 
er, broader bill. As a general pattern, Schnell 
(1994) noted that Common Black-Hawks of conti- 
nental (inland) North and Central America are 
largest. Mainland anthracinus populations inhabit- 
ing mangrove habitat tend to be smaller and 
browner than others. The race B. subtilis rhizopho- 
rae, which inhabits mangrove habitat (Monroe 
1963, 1968, Blake 1977), shows a dark-brown plum- 
age. Our observations revealed that Cuban birds, 
also mangrove inhabitants, are consistently brown- 
er with substantial differences in plumage pattern 
compared with mainland birds. Thus, such color 
differences may be a result of ecological parallel- 
ism, rather than of phylogenetic relationships. 

The species of Buteogallus are partial to wetlands, 
swampy woods, and seacoasts (Amadou 1982). In 
its mainland range, anthracinus has been charac- 
terized as inhabiting woodlands around coastal 


December 2005 


Cuban Black-Hawk Status and Biology 


361 


swamps, ponds, and streams, and especially man- 
groves in the swampy woodlands adjacent to the 
poorly-drained inlands that are affected by tide- 
waters (Phillips et al. 1964, Wetmore 1965, Davis 
1972, Schnell 1994). Wetmore (1965) noted that 
along large rivers they extend their range farther 
inland. Thomas (1908) reported anthracinus in 
stretches of sand dunes and savannas with clumps 
of palmettos and pines. The Cuban population 
shows a similar preference for lowland coastal ar- 
eas. Gundlach (1893) and Bangs (1905) noted 
gundlachii was found only in mangrove swamps and 
on the banks of large rivers. In broad contrast, the 
other West Indian population, Buteogallus anthraci- 
nus cancrivorus of St. Vincent, mainly keeps to the 
high wooded valleys, although it seldom occurs far 
from water (Lister 1880, Clark 1905b, Bond 
1956a). 

Cuban populations of the black-hawk breed 
from January through June (Garrido and Kirkcon- 
nell 2000) , with egg-laying occurring in late March 
or April. Bangs (1905) collected a female contain- 
ing a soft-shelled egg and found another tending 
a nest on 15 April. Bond (1950) reported a nest 
with a newly-hatched chick on 4 April. Garrido and 
Schwartz (1969) and Valdes Miro (1984) com- 
mented gundlachii builds its nest at a considerable 
distance above the ground. Gundlach (1876) re- 
ported a nest at 8 “varas” (6.8 m), whereas Bond 
(1936) noted one at 6.2 m. 

Nests of the Cuban form are typically rough 
structures of twigs, lined with green leaves and, 
sometimes, debris (Gundlach 1893, Bond 1936, 
Garrido and Schwartz 1969, Valdes Miro 1984). 
Bond (1936), describing nests found in St. Vincent 
{B. a. cancrivorus) and Cuba {gundlachii), noted, 
“The nest, a rough mat of sticks, is placed at vari- 
ous elevations in trees.” All nests located by us at 
Los Indios in 1996-98 were in mangroves {Avicen- 
nia, Rhizophora). In contrast, Bond (1936) de- 
scribed black-hawk nests in St. Vincent as “placed 
on top of clumps of mistletoe and were rather 
small.” As Bond (1936) suggested, nests of the Cu- 
ban species are somewhat larger than those of 
birds in St. Vincent. Schnell (1994) gave the di- 
mensions of mainland anthracinus nests as ranging 
from 38 cm diameter X 20 cm deep to 1.2 m di- 
ameter X 0.67-1.2 m deep. Bangs (1905) and 
Bond (1936) also noted gundlachii re-used nests in 
more than one season, which we believe accounts, 
in part, for the larger nest size of Cuban birds. 

Black-hawks at Los Indios were remarkably non- 


aggressive toward humans at their nests and al- 
lowed us to approach much closer than other local 
raptor species tolerated, perhaps relying on their 
cryptic behavior to avoid detection at the nest. 
Others have also noted this tolerance in Cuban 
black-hawks (Todd 1916, Barbour 1923, Garrido 
and Schwartz 1969). 

Schnell (1994) reviewed available egg specimens 
for Buteogallus anthracinus, summarizing mean 
measurements from Bent (1937) as 57.3 X 44.9 
mm {N = 60 eggs) and examples in the Western 
Foundation of Vertebrate Zoology as 57.30 X 45.50 
mm (V = 12 clutches, 19 eggs; range = 52.61- 
62.02 mm length, 42.69-47.35 mm breadth). Eggs 
of anthracinus we measured at the Delaware Mu- 
seum of Natural History averaged 57.46 (53.1- 
63.2) X 45.25 (41.7-49.1) mm {N = 13 clutches, 
21 eggs). Interestingly, an egg reported from St. 
Vincent is at the high end for the species: 61 X 47 
mm (Bond 1936) and exceeds the range for gund- 
lachii. Eggs of gundlachii we measured at Los Indios 
averaged only slightly smaller than those of main- 
land B. anthracinus analyzed by Schnell (1994). 
Gundlach (1876) reported that Cuban eggs mea- 
sured 58 X 45 mm, whereas Bangs (1905) reported 
56 X 45.5 mm. Measurements presented by Valdes 
Miro (1984) are obviously in error; i.e., x = 56.0 
(range = 55.0-57.0) X 24.6 (23.0-26.5) mm. The 
mean mass (61.0 ± 1.8 g) of eggs we measured at 
Los Indios was somewhat lighter compared with 
SchnelFs estimated mean mass of 63.8 g for anth- 
racinus. 

Although we observed differences in egg color- 
ation and pattern among anthracinus, subtilis, and 
gundlachii, these characters show considerable var- 
iation and do not appear to be a good character 
for determining relationships (L. Kiff pers. 
comm.). 

Schnell (1994) noted that, in general, clutch size 
of Buteogallus anthracinus decreased from two eggs 
in the northern range to one in the southern 
range; several reported three-egg clutches were 
questionable. Clutch sizes at Los Indios fell within 
that range, averaging 1.57 eggs per clutch. 

Buteogallus anthracinus feeds mainly on inver- 
tebrates and lower vertebrates, with occasional 
small birds or mammals in the diet (Schnell 1994). 
Eor mainland populations, Thomas (1908) report- 
ed anthracinus preying on burrowing land crabs, 
which form almost the sole diet of the hawks in 
British Honduras (Belize). The St. Vincent popu- 
lation {B. a. cancrivorus) reportedly feeds on cray- 


362 


Wiley AND Garrido 


VoL. 39, No. 4 


fish and freshwater crabs (Lister 1880). In Cuba, 
Gundlach (1893) reported remains of crustaceans, 
as well as frogs, snakes, and fishes in the stomachs 
of black-hawks. Barbour (1943) reported land crabs 
as its prey in Cuba. Garrido and Kirkconnell 
(2000) reported its prey as mainly crabs and birds, 
whereas Ramsden (C. Ramsden, Museo de His- 
toria Natural, Universidad de Oriente, Santiago de 
Cuba unpubl. data) noted the hawk fed on crabs 
and fishes. 

The hunting behavior of Buteogallus, in general, 
has been characterized as sluggish. Schnell {in 
Palmer 1988, 1994) noted B. anthracinus normally 
hunts from a stationary perch, often near the 
ground, from branches up to 15 m high, on boul- 
ders, other low perches, and gravel beds along 
streams. For Cuban hawks, Barbour (1923) de- 
scribed crab predation similar to our observations: 
“The hawk pounces on the crab, gathers the legs 
and claws of each side in one of its feet, and reach- 
ing down removes the carapace by hooking the bill 
under its front edge.” Kirkconnell and Garrido 
(1991) reported gundlachii drowning its avian prey 
(Common Moorhen [Gallinula chloropus]), which 
they suggested was unusual and perhaps related to 
the abundant rain that caused the raising of the 
water level in the swamp, rendering crabs difficult 
to find. 

We observed Cuban Black-Hawks caching crab 
prey near their nest, a habit that has also been re- 
ported for B. anthracinus in mainland sites (Thom- 
as 1908, Schnell 1991, 1994). 

As noted by Schnell (1994), descriptions of the 
vocal behavior of Buteogallus anthracinus have been 
confusing and conflicting. Schnell (1994) charac- 
terized the common call (= alarm call) as of a 
complex, un-raptor-like quality. The common call 
of mainland Buteogallus anthracinus is distinct from 
the three-note call of gundlachii, consisting of 9-24 
notes (Reynard and Garrido 1988, Schnell 1994) 
(Figs. 2A, 2C— F, Table 5). Similarly, the common 
call of B. subtilis is distinct from that of gundlachii, 
consisting of several, rapidly-repeated elements, 
described by Ferguson-Lees and Christie (2001) as 
a series of shrill whistles, indistinguishable from 
that of anthracinus. The call of Buteogallus aequinoc- 
tialis is a series of six or seven whistle-like notes, 
the first three rapid, followed by slower and de- 
scending elements (Fig. 2H; Ferguson-Lees and 
Christie 2001). Finally, B. meriodionalis has a call 
consisting of a prolonged whistle, described as 


“eeeeee-eh” or ''kree-ee-ee-er” (Ferguson-Lees and 
Christie 2001). 

Conclusions 

We consider Buteogallus anthracinus (with its geo- 
graphical races, cancrivorus and anthracinus), B. 
urubitinga, B. aequinoctialis, and B. gundlachii as sep- 
arate species. This treatment of the Cuban popu- 
lation agrees with Wetmore (1965:234), who stated 
the other forms stand apart: “. . . from the bird of 
the island of Cuba which it appears appropriate to 
treat as a separate species, Buteogallus gundlachii.” 
Thus, the Cuban Black-Hawk Buteogallus gundlachii 
Cabanis, 1854 (1855), becomes a species endemic 
to Cuba, distributed in the main island, where it is 
relatively uncommon and quite localized, Isla de 
Pinos, and many of the keys of the Cuban Archi- 
pelago. 

Acknowledgments 

A grant from the American Museum of Natural History 
allowed Garrido to undertake studies of West Indian 
birds in its and other United States Museums. We grate- 
fully acknowledge help and support from the RARE Cen- 
ter for Tropical Conservation, through John Guarnaccia 
and Victor L. Gonzalez, enabling visits to collections in 
several museums. We also thank the curators of the 
American Museum of Natural History, the Smithsonian 
Institution, the Museum of Comparative Zoology, Loui- 
siana State University Museum of Natural History, Acad- 
emy of Natural Science, Delaware Museum of Natural 
History, Instituto de Ecologia y Sistematica, and Museo 
Nacional de Historia Natural de Cuba for their cooper- 
ation. Wiley’s research in Isla de Pinos was supported by 
grants from Wildlife Preservation Trust International, 
World Parrot Fund, International Crane Foundation, and 
the U. S. Department of the Interior. We are particularly 
grateful to Lie. Xiomara Galvez, Empresa Nacional para 
la Conservacion de la Flora y Fauna, for logistical support 
and field assistance at Los Indios. For use of tape record- 
ings of Buteogallus vocalizations, we thank George B. Rey- 
nard, Paul Schwartz, L. Irby Davis, R. Grimshaw, J 
Schnell, Richard Ranft, and the British Library Sound 
Archive. We thank Lloyd F. Kiff, Jay H. Schnell, and an 
anonymous reviewer for their helpful suggestions, which 
greatly improved the manuscript. 

Literature Cited 

Alain, H. 1946. Notas taxonomicas y ecologicas sobre la 
flora de Isla de Pinos. Contr. Ocas. Museo Hist. Nat. 
Coleg. La Salle 7:11-115. 

Aldrich, J.W. and B.P. Bole, Jr. 1937. The birds and 
mammals of the western slope of the Azuero Penin- 
sula, Republic of Panama. Sci. Publ. Cleveland Mus 
Nat. Hist. 7:1-196. 

Amadon, D. 1961. Remarks on the genus Buteogallus. Nov 
Colombianas 1:358-360. 


December 2005 


Cuban Biack-Hawk Status and Biology 


363 


. 1982. A revision of the sul>Buteonine hawks (Ac- 

cipitridae, Aves). Am. Mus. Nov. No. 2741:1-20. 

American Ornithologists’ Union. 1998. Check-list of 
North American birds, 7th Ed. American Ornitholo- 
gists’ Union, Washington, DC U.S.A. 

Bangs, O. 1905. The Cuban Crab Hawk, Urubitinga gund- 
lachii (Cabanis). Auk 22:307-309. 

and W.R. Zappey. 1905. Birds of the Isle of Pines. 

Am. Nat. 39:179-215. 

Barbour, T. 1923. The birds of Cuba. Mem. Nuttall Or- 
nithol. Club 6. 

. 1943. Cuban ornithology. Mem. Nuttall Omithol. 

Club 9. 

Bent, A.C. 1937. Life histories of North American birds 
of prey. Part I. U.S. Nat. Mus. Bull. 167. 

Blake, E.R. 1977. Manual of Neotropical birds. Vol. I. 
University Chicago Press, Chicago, IL U.S.A. 

Bond, J. 1936. Birds of the West Indies. Acad. Nat. Sci. 
Philadelphia, PA U.S.A. 

. 1950. Check-list of birds of the West Indies. Acad. 

Nat. Sci. Philadelphia, PA U.S.A. 

. 1956a. Check-list of birds of the West Indies, 4th 

Ed. Acad. Nat. Sci. Philadelphia, PA U.S.A. 

. 1956b. First supplement to the Check-list of birds 

of the West Indies (1956). Acad. Nat. Sci. Philadel- 
phia, PA U.S.A. 

Brown, L. and D. Amadon. 1968. Eagles, hawks and fal- 
cons of the world. Vol. 2. McGraw-Hill Book Co., New 
York, NYU.S.A. 

Cabanis, J.L. 1855. Dr. J. Gundlach’s Beitrage zur Ornit- 
hologie Cuba’s. J. Ornithol. 2 (extra-heft 1854 [pub- 
lished 1855]):77-87. 

Clark, A.H. 1905a. The Crab Hawk {Urubitinga) in the 
island of St. Lucia, West Indies. Auk 22:210. 

. 1905b. Preliminary descriptions of three new 

birds from St. Vincent, West Indies. Proc. Biol. Soc. 
Wash. 18:61-63. 

. 1905c. Birds of the southern Lesser Antilles. Proc. 

Boston Soc. Nat. Hist. 32:203-312. 

Cory, C.B. 1887. The birds of the West Indies, including 
the Bahama Islands, the Greater and the Lesser An- 
tilles, excepting the islands of Tobago and Trinidad. 
ArA 4:37-51. 

. 1892. Catalogue of West Indian birds, containing 

a list of all species known to occur in the Bahama 
Islands, the Greater Antilles, the Caymans, and the 
Lesser Antilles, excepting the islands of Tobago and 
Trinidad, Privately published by the author, Boston, 
MA U.S.A. 

Davis, L.L 1972. A field guide to the birds of Mexico and 
Central America. University Texas Press, Austin, TX 
U.S.A. 

Dickinson, E.C. [Ed.]. 2003. The Howard and Moore 
complete checklist of the birds of the world, rev. and 
enlarged 3rd Ed. Princeton University Press, Prince- 
ton, NJ U.S.A. 


Evans, P.G.H. 1990. Birds of the eastern Caribbean. Mac- 
Millan Press Ltd., London, England. 

Ferguson-Lees, j. and D.A. Christie. 2001. Raptors of 
the world. Houghton Mifflin Co., New York, NY 
U.S.A. 

Friedmann, H. 1950. Birds of North and Middle America, 
Falconiformes. U.S. Nat. Mus. Bull. 50, pt. 11. 

Garrido, O.H. and a. Kirkconnell. 2000. Field guide to 
the birds of Cuba. Cornell University Press, Ithaca, NY 
U.S.A. 

and a. Schwartz. 1969. Anfibios, reptiles y aves 

de Cayo Candles. Poeyana 67:44. 

Gundlach, j. 1854. Beitrage zur Ornithologie Cuba’s. 
Nach Mittheilungen des Reisenden an Hr. Bez.-Dir. 
Sezekorn in Cassel; von Letzterem zusammengestellt. 
Mit Zustzen und Anmerkungen geordnet vom He- 
rausgeber. /. Ornithol. 2 (extra-heft) :77-87. 

. 1865-1 866a. Revista y catalogo de las aves Cu- 

banas. Pages 165-180 in Repertorio Fisico Natural de 
la Isla de Cuba. Vol. 1 (F. Poey, [Ed.]). Imprenta del 
Gobierno y Capitania General, La Habana. 

. 1865-1 866b. Revista y catalogo de las aves Cu- 

banas. Pages 221-224 in Repertorio Fisico Natural de 
la Isla de Cuba. Vol. 1 (F. Poey, [Ed.]). Imprenta del 
Gobierno y Capitania General, La Habana. 

. 1871. Neue Beitrage zur Ornithologie Cubas. 

Nach eigenen 30 jahrigen Beobachtungen zusam- 
mengestellt./. Omithol. 19:353-378. 

. 1876. Contribucion a la ornitologia Cubana. La 

Habana: “La Antilla.’’ 

. 1893. Ornitologia Cubana. La Habana: La Mo- 

derna. 

Gurney, J.H. 1876. Notes on a “Catalogue of the Acci- 
pitres in the British Museum,” by R. Bowdler Sharpe 
(1874). Ibis, series 3, 6:467—493. 

. 1934. A list of the diurnal birds of prey, with 

references and annotations. John Van Voorst, Lon- 
don, England. 

Hellmayr, C., and H. Conover. 1949. Catalogue of birds 
of the Americas. Field Mus. Nat. Hist. Zool. Ser., 13, part 
1(4). 

Jennings, O.E. 1917. A contribution to the botany of the 
Isle of Pines, Cuba, based upon the specimens of 
plants from that island contained in the herbarium of 
the Carnegie Museum under date of October, 1916. 
Ann. Carnegie Mus. 11:19—290. 

Kirkconnell, A. and O. Garrido. 1991. Aberrante com- 
portamiento de caza del Gavilan Batista Buteogallus 
anthracinus gundlachii (Aves: Accipitridae) en Cuba 
Vol. Migrat. No. 16:27-28. 

Kleinbaum, D.G. and L.L. Kupper. 1978. Applied regres- 
sion analysis and other multivariable methods. Dux- 
bury Press, North Scituate, MA U.S.A. 

Lister, C.E. 1880. Field-notes on the birds of St. Vincent, 
West Indies. Ibis, ser. 4, 4:38-44. 

Mayr, E. and G.W. Cottrell [Eds.]. 1979. Check-list of 
birds of the world. Vol. 1, 2nd Ed. (rev. of the work 


364 


Wiley AND Garrido 


VoL. 39, No. 4 


of Peters, J.L.). Mus. Comp. Zool. Annu. Rep., Cam- 
bridge, MA U.S.A. 

AND L.L. Short. 1970. Species taxa of North 

American birds. Publ. Nuttall Ornithol. Club No. 9. 

Monroe, B.L., Jr. 1963. Three new subspecies of birds 
from Honduras. Occas. Pap. Mus. Zool., Louisiana State 
University 136. 

. 1968. A distributional survey of the birds of Hon- 
duras. American Ornithologists’ Union Ornithol. Mo- 
nogr. No. 7. Washington, DC U.S.A. 

Palmer, R.S. [Ed.]. 1988. Handbook of North American 
birds. Vol. 4. Yale Universtiy Press, New Haven, CT 
U.S.A. 

Peters, J.L. 1931. Check-list of birds of the world. Vol. 1. 
Harvard University Press, Cambridge, MA U.S.A. 

Phillips, A., J. Marshall, and G. Monson. 1964. The 
birds of Arizona. University Arizona Press, Tucson, AZ 
U.S.A. 

Reynard, G.B. and O.H. Garrido. 1988. Bird songs in 
Cuba/Cantos de aves en Cuba. Cornell Lab. Ornit- 
hol., Ithaca, NY U.S.A. 

Ridgely, R.S. AND PJ- Greenfield. 2001. The birds of Ec- 
uador: status, distribution and taxonomy. Christopher 
Helm, London, England. 

Ridgway, R. 1876. Studies of the American Falconidae. 
Dept. Interior, U. S. Geolog. Geograph. Survey Terri- 
tories 2:91-182. 

Schnell, J.H. 1991. A closer look: Common Black-Hawk. 
Birding 23:282-285. 

. 1994. Common Black-Hawk, Buteogallus anthraci- 

nus. In Poole, A.F., and F.B. Gill [Eds.] The birds of 
North America, No. 122. Acad. Nat. Sci. Philadelphia, 
and American Ornithologists’ Union, Washington, 
DC U.S.A. 

Sharpe, R.B. 1874. Catalogue of the birds in the British 
Museum. Vol. I. Catalogue of the Accipitres, or diur- 


nal birds of prey, in the collection of the British Mu- 
seum. British Museum, London, England. 

. 1899. A hand-list of the genera and species of 

birds. Vol. I. British Museum, London, England. 

Sibley, C.G. and B.L. Monroe, Jr- 1990. Distribution and 
taxonomy of birds of the world. Yale University Press, 
New Haven, CT U.S.A. 

Snyder, N.F.R. and J.W. Wiley. 1976. Sexual size dimor- 
phism in hawks and owls in North America. American 
Ornithologists’ Union Ornithol. Monogr. No. 20. 

SPSS. 1999. SPSS for Windows 10.0.5. SPSS Inc., Chicago, 
IL U.S.A. 

Stiles, EG., and A.F. Skutch. 1989. A guide to the birds 
of Costa Rica. Comstock Pub. Assoc., Ithaca, NY 

U.S.A. 

Storer, R.W. 1966. Sexual dimorphism and food habits 
in three North American accipiters. Auk 83:423-436. 

Swann, H.K. 1921-1922. A synopsis of the Accipitres (di- 
urnal birds of prey) comprising species and subspe- 
cies described up to 1920, with their characters and 
distribution. 2*“^ edition, revised and corrected 
throughout. Wheldon and Wesley, London, England. 

. 1930. A monograph of the birds of prey (Order 

Accipitres). Vol. I. Part VIII. Wheldon and Wesley, 
London, England. 

Thomas, G.B. 1908. The Mexican Black Hawk. Condor 10' 
116-118. 

Todd, W.E.C. 1916. The birds of the Isle of Pines. Incor- 
porating the substance of the field-notes by Gustav A 
Link. Ann. Carnegie Mus. 10:146-296. 

Valdes Miro, V. 1984. Datos de nidificacion sobre las 
aves que crian en Cuba. Poeyana 282:1-27. 

Wetmore, a. 1965. The birds of the Republic of Panama, 
pt. 1. Smithson. Misc. Collect. 150. 

Received 28 June 2004; accepted 28 July 2005 

Associate Editor: Cheryl R. Dykstra 


J. Raptor Res. 39(4):365-377 
© 2005 The Raptor Research Foundation, Inc. 


HOME RANGE AND HABITAT USE OF NORTHERN SPOTTED 
OWLS ON THE OLYMPIC PENINSULA, WASHINGTON 

Eric D. Forsman,i TimmothyJ. Kaminski,^ Jeffery C. Lewis,^ Kevin J. Maurice,^ and 

Stan G. Sovern^ 

USD A Forest Service, Pacific Northwest Research Station, 3200 Jefferson Way, Corvallis, OR 97331 US. A. 

Cheron Ferland and Elizabeth M. Glenn 

Department of Fisheries and Wildlife, Oregon State University, Corvallis, OR 97331 US. A. 

Abstract. — ^We studied movements and habitat selection of 20 adult northern Spotted Owls (Strix oc- 
cidentalis caurina) on the Olympic Peninsula, Washington in 1987-89. Median annual home range size 
of individual owls was 1147 ha based on the 75% isopleth of the Fixed Kernel (FK), 2406 ha based on 
the 95% FK, and 2290 ha based on the 100% Minimum Convex Polygon (MCP). Annual ranges of 
individual owls tracked >1 yr overlapped by a mean of 70-73%, depending on which estimator was 
used. Size of annual and cumulative ranges was negatively correlated with the amount of old forest 
within the cumulative MCP home range and within a 4.3 km radius of the center of activity. Overlap of 
annual ranges of owls that were paired averaged 64 ± 5% based on the MCP and 69 ± 5% based on 
the 95% FK. On average, ranges used during the nonbreeding season overlapped breeding season 
ranges by 65.0 ± 4.5%, and breeding season ranges overlapped nonbreeding season ranges by 62.6 ± 
4.9%. Compositional analysis of habitat selection indicated that old forests were the most preferred 
cover type for foraging and roosting and that dear-cuts and non-forest cover types were rarely used. 
There was little evidence that owls selected riparian areas or forest edges for foraging or roosting. Our 
observations are consistent with the hypotheses that northern Spotted Owls use larger foraging areas 
in regions where northern flying squirrels {Glaucomys sabnnus)arG their primary source of food, that 
they prefer old forests for foraging and roosting, and that their home ranges become larger as the 
amount of old forest declines. The large size of annual ranges on the Olympic Peninsula may be a 
response to low prey biomass. 

Key Words: Northern Spotted Owl, Strix occidentalis caurina; home range, habitat use, radiotelemetry, Olympic 
Peninsular, Washington. 


RANGO DE HOGAR Y USO DE habitat DE strix occidentalis CAURINA EN OLYMPIC PEN- 
INSULA, WASHINGTON 

Resumen. — Estudiamos los movimientos y la seleccion de habitat de 20 individuos adultos de Strix occi- 
dentalis caurina en Olympic Peninsula, Washington, entre 1987 y 1989. La mediana del tamaho del area 
de hogar de un individuo fue de 1147 ha basada en la isolinea de 75% del kernel fijo (KE), 2406 ha 
basada en el 95% KF y 2290 ha basada en el 100% del poligono convexo minimo (PCM). Los rangos 
anuales de los individuos seguidos por menos de un ano se superpusieron en promedio entre un 70% 
y un 73%, dependiendo del estimador que usamos. Los tamanos de los rangos anuales y acumulativos 
se correlacionaron negativamente con la cantidad de bosque maduro presente dentro del PCM acu- 
mulativo del rango de hogar y a menos de 4.3 km del centre de actividad. La superposicion promedio 
de los rangos de hogar anuales de individuos que confer maban parejas fue de 64 ± 5% basado en el 
PCM y 69 ± 5% basado en el 95% del KF. En promedio, los rangos usados durante el periodo no 
reproductive se superpusieron con los rangos del periodo reproductive en 65.0 ± 4.5%, y los rangos 
del periodo reproductive se superpusieron con los rangos del periodo no reproductive en 62.6 ± 4.9%. 


^ Email address: eforsman@fs.fed.us 

^Present address: U.S. Forest Service, Bridger-Teton National Forest, P.O. Box 1888, Jackson, WY 83001 U.S.A. 

^ Present address: Washington Department of Fish and Wildlife, 600 Capitol Way North, Olympia, WA 98501 U.S.A. 

Present address: U.S. Fish and Wildlife Service, Portland Field Office, 2600 SE 98*, Portland, OR 97266 U.S.A. 

’’ Present address: Department of Fisheries and Wildlife, Oregon State University, Corvallis, OR 97331 U.S.A. 


365 


366 


Forsman et al. 


VoL. 39, No. 4 


Los analisis de composicion de los ambientes seleccionados indicaron que los bosques maduros fueron 
el dpo de cobertura preferida para alimentarse y reposar, mientras que las areas completamente taladas 
y no boscosas fueron usadas en muy pocas ocasiones. Encontramos muy poca evidencia de que las 
lechuzas seleccionan las areas riparias o los hordes de bosque para alimentarse o reposar. Nuestras 
observaciones son consistentes con la hipotesis de que S. o. caurina usa grandes areas de forrajeo en las 
regiones donde las ardillas voladoras {Glaucomys sabrinus) son su fuente principal de alimento, que 
prefieren bosques maduros para alimentarse y reposar y que sus areas de hogar aumentan a medida 
que disminuye la cantidad de bosque maduro. El gran tamano de los rangos anuales en Olympic Pen- 
insula podria responder a una baja biomasa de presas. 

[Traduccion del equipo editorial] 


Spotted Owls {Strix occidentalis) exhibit consid- 
erable variation in home range size and patterns 
of seasonal movements, both within and among re- 
gions. For example, in some parts of their range, 
Spotted Owls may migrate during winter, moving 
16-58 km from their breeding season ranges into 
lowland forests (Laymon 1989, Zabel et al. 1992). 
In other regions, they are largely resident in the 
same areas throughout the year (Forsman et al. 
1984, Carey et al. 1990, 1992). 

Home ranges and habitat selection of Spotted 
Owls have been studied extensively in Oregon and 
California, but with the exception of a study by 
Hamer (1988), little information is available from 
Washington. We examined home ranges and hab- 
itat selection of northern Spotted Owls on the 
Olympic Peninsula, Washington to determine if 
patterns of habitat use differed near the northern 
edge of the range of the owl compared to earlier 
studies conducted in Oregon (e.g., Forsman et al. 
1984, Carey et al. 1990, 1992, Carey and Peeler 
1995) and northern California (Solis and Gutier- 
rez 1990, Zabel et al. 1992, 1995). 

Study Area 

We conducted our study on two areas on the west side 
of the Olympic Peninsula, one located 3 km SE of the 
town of Forks, Clallam County, and the other located 10 
km SE of the town of Quinault, Jefferson County (Fig. 
1). Both areas were located on the Olympic National For- 
est, had similar climate, topography and vegetation, and 
will hereafter be referred to collectively as the “study 
area.” 

The study area was characterized by mountainous ter- 
rain covered by forests of western hemlock (Tsuga hete- 
rophylla) and western redcedar {Thuja plicata). Sitka 
spruce {Picea sitchensis) was common on mesic, low ele- 
vation areas, and Douglas-fir {Pseudotsuga menziesii) and 
Pacific silver fir (Abies amabilis) were often intermixed 
with western hemlock on upland sites (Henderson et al. 
1986). Elevations ranged from 150-1500 m. Precipitation 
ranged from 280-460 cm/yr, mostly falling as rain during 
October-May. 

The area included a mosaic of serai stages, ranging 
from clearings in which all trees had been recently har- 


vested (clear-cuts) to old-growth forests in which oversto- 
ry trees were over 500 yr old (Henderson et al. 1986). 
Approximately half of the area had been clear-cut within 
the previous 30 yr, but harvested areas were not uniform- 
ly distributed within the study area. Some areas were 
heavily fragmented by recent clear-cuts, whereas other ar- 
eas had extensive blocks of mature and old-growth forest. 
Much of the study area was hit by hurricane-force winds 
in 1921 which severely damaged many stands (Pierce 
1921). As a result, many stands included a mixture of 60- 
80-yr-old trees that regenerated after the wind event, in- 
terspersed with old trees (80-500+ yr) that survived the 
windstorm. All types of natural (unlogged) forest typical- 
ly had high canopy closure (65-80%), high variation in 
tree size and age, and high volumes of logs and snags 
(Henderson et al. 1986). Regenerating stands of young 
trees in clear-cut areas were usually even-aged, with high 
canopy closure. 

Methods 

Capture and Radio-marking. We captured owls with 
noose poles (Forsman 1983) and marked them with 
back-pack transmitters (Model P2, AVM Instrument 
Company, Livermore, CA U.S.A.), as described by Fors- 
man et al. (1984). Total mass of transmitter and harness 
was 18-20 g, and transmitter life was 9-15 mo. We tried 
to obtain a minimum of 12 mo of data from each owl. 
We replaced transmitters on six individuals after 9-12 
mo, and tracked them for nearly 2 yr. 

Sampling Schedule. We attempted to obtain one noc- 
turnal foraging location per night on each owl at least 3 
nights per wk, and one diurnal roost location per owl at 
least 3 d per wk. Our sampling schedule was intended to 
reduce autocorrelation between sequential locations 
(Swihart and Slade 1985a, 1985b). However, Aebischer et 
al. (1993) and Otis and White (1999) have suggested that 
autocorrelation is generally irrelevant when individual 
animals are used as the sample unit in home-range stud- 
ies, so we used all of our data, including a few cases (129 
of 7346 locations) when we obtained 2-3 locations on 
the same owl in one night. We classified all locations as 
foraging locations if they occurred from 0.5 hr after sun- 
set to 0.5 hr before sunrise. We excluded locations of 
incubating or brooding females from analyses of habitat 
selection, until females began to forage when the young 
were about 2 wk old. 

Radio Triangulation. We estimated owl locations by tri- 
angulating with a Telonics hand-held H-antenna and TR2 
receiver (Telonics, Mesa, AZ U.S.A.). We used a hand- 
held compass to estimate azimuths from ^3 locations 


December 2005 


Spotted Owl Habitat Use 


367 



Figure 1. Location of radiotelemetry study areas on the Olympic Peninsula, Washington, 1987-89. 


) 

1 

I 


along roads (Guetterman et al. 1991). Azimuths were 
plotted on 1:12 000 or 1:24000 scale U.S. Geological Sur- 
vey orthophotos or topographic maps. We considered the 
position of the owl to be the geometric center of the 
polygon formed by the intersection of ^3 bearings 
(Nams and Boutin 1991). If weak signals or inconsisten- 
cies in the direction of bearings caused us to suspect sig- 
nal deflection or movement of an owl during triangula- 
tion, we discarded the location. We used all locations to 
estimate home ranges, but only locations with error poly- 
gons ^8 ha were used for analyses of habitat use. 

Telemetry Error. We estimated telemetry error with 63 
blind trials in which one observer placed transmitters in 
trees in owl home ranges and another observer then tri- 
angulated on the transmitters at night. The median dis- 
tance between estimated and actual transmitter locations 
was 100 m (x = 140 ± 17 m). This estimate was similar 
to or less than error estimates in previous telemetry stud- 
ies of Spotted Owls (Carey et al. 1990, Glenn et al. 2004). 
Errors of this magnitude undoubtedly resulted in some 
locations falling in the wrong cover types, but we made 
the assumption that classification errors due to telemetry 
error were similar in all cover types, and that our overall 
assessment should reflect actual habitat use. 


Home-range Estimation. We estimated cumulative and 
annual ranges with the Minimum Convex Polygon 
(MCP) and Fixed-Kernel (FK) methods (Hayne 1949, 
Seaman and Powell 1996). For estimates of MCP ranges, 
we used 100% MCP polygons. For FK estimates, we used 
95% and 75% isopleths, which we interpreted as the 
“home range,” and “area of concentrated use,” respec- 
tively. We used Program CALHOME (Kie et al. 1996) to 
estimate MCP ranges and Version 4.28 of Program ICER- 
NELHR (Seaman et al. 1998) to estimate FK ranges. Con- 
trary to the recommendation of Seaman and Powell 
(1996), we used the FK method without least-squares- 
cross-validation (LSCV). We did so because we believe 
that kernel estimates based on locations where owls stop 
long enough for the observer to obtain a location tend 
to underestimate home range areas of owls (because 
movements across intervening non-forest areas usually 
happen so quickly that they cannot be documented with 
a point on the map) . Thus, we feel that the LSCV option, 
which tends to fit the home range isopleth more tightly 
to the observed points, is likely to cause an even greater 
underestimate of home ranges. We used all locations for 
MCP estimates, but we only used foraging locations for 
FK estimates (because FK estimates that include large 


368 


Forsman et al. 


VoL. 39, No. 4 


Table 1. Vegetation cover types used to map landscapes for analyses of habitat use by northern Spotted Owls on 
the Olympic Peninsula, Washington, 1987-89. 


Old Forest: Multilayered stands of western hemlock and western redcedar in which the dominant overstory trees 
were typically ^100 cm DBH. Pacific silver fir was often subdominant or codominant with hemlock or redcedar. 
Douglas-fir was codominant on a few areas. Also included mixed-age stands of mature and old forest in which 
both age classes were common. Many of the latter stands were the result of a hurricane force windstorm in 
January 1921 (Pierce 1921). 

Mature Forest: Conifer-dominated stands in which the overstory trees were typically 50—99 cm DBH. 

Young Forest: Relatively even-aged stands in which most trees were 31-60 cm DBH. Regenerated on burned areas 
and old dear-cuts. 

Mixed-young Forest: Same as Young Forest except with inclusions of mature trees, usually remnants left during 
previous fires or harvest. 

Pole-sapling: Single-layered conifer stands in which most trees were 10-30 cm DBH. Mostly young stands regenerat- 
ing on old dear-cuts. 

Hardwood/Riparian: Riparian areas dominated by red alder {Alnus rubra), bigleaf maple {Acer macrophyllum) , and 
variable amounts of western redcedar. 

Clear-cut/Non-forest: Recent dear-cuts dominated by bare soil, grasses, shrubs or small seedling conifers. Also in- 
cluded small areas of meadows, gravel pits, and agricultural, or residential areas. 


numbers of roosting locations clustered at the nest site 
or central place will underestimate foraging areas during 
the breeding season) . 

Estimation of Annual, Cumulative, and Seasonal Rang- 
es, Although we marked some owls in June or July of 

1987, we did not begin regular sampling of most individ- 
uals until late July or August 1987, For these owls, we 
estimated the first annual range through the end of July 

1988. If they were monitored after July 1988, we com- 
puted a second annual range for the second year. A few 
individuals were not marked until fall 1987 or summer 
1988, in which case the annual range was estimated for 
one year only. There was only a weak positive correlation 
between the number of days in the tracking period and 
estimates of annual home-range size, regardless of which 
home range estimator was used (95% FK r^^ ~ 0-221, P 
= 0.223; MCP r^^ = 0.208, P = 0.253). Therefore, we 
used all annual ranges for comparisons among owls, re- 
gardless of the monitoring period. 

For six owls tracked in both years, we estimated the 
cumulative range from the union of the annual ranges 
(range A -I- range B minus the area of overlap). Estimates 
of home range overlap between years, seasons, pair mem- 
bers, or owls on adjacent territories were based on the 
percent of range A overlaid by range B or the percent of 
range B overlaid by range A. In most cases, we computed 
overlap of ranges based on three different frames of ref- 
erence (75% FK, 95% FK, and 100% MCP). For estimates 
of overlap of seasonal ranges, we only used the 95% FK. 

For seasonal analysis of home ranges, we divided each 
year into two phonological periods, the “breeding sea- 
son” (March— August) , when Spotted Owls nest and feed 
young, and the “nonbreeding season” (September— Feb- 
ruary) , when Spotted Owls are largely solitary. Estimates 
of seasonal ranges were limited to owls tracked sl20 d 
during the season of interest. 

Habitat Mapping and Assessment of Habitat Use. We 
examined second-order habitat selection (i.e., use of dif- 
ferent forest cover types within the home range of each 


owl). We developed a cover-type map of the study area 
that included seven cover types based on structural dif- 
ferences in vegetation as determined from on-the-ground 
examination of stands and aerial photo interpretation 
(Table 1). We visited virtually all stands within the study 
area on one or more occasions to determine the size and 
species composition of trees. We did not use canopy clo- 
sure to differentiate among cover types because nearly all 
forests on the study area had relatively high (>70%) can- 
opy closure, regardless of stand age or tree size. Cover 
types were mapped on 1:12 000 scale orthophotos and 
digitized into an ARC/INFO (ESRI Inc., Redlands, CA 
U.S.A.) GIS layer. For convenience, we use the term cover 
type, even though we recognize that our designation of 
cover type was based on only one component of habitat 
(i.e., vegetation structure). Site visits to 403 randomly se- 
lected grid coordinates indicated tbat map accuracy was 
83%. 

We used compositional analysis (Aebischer et al. 1993) 
to evaluate relative preference of cover types for foraging 
and roosting. This method treats the individual as the 
sample unit, accounts for lack of independence among 
proportions, is not sensitive to serial correlation between 
locations, and is based on a unique set of observed and 
expected values for each cover type in the home range 
of each individual. Expected use was equal to the pro- 
portion of the cumulative MCP home range covered by 
each cover type, and the observed use was the proportion 
of locations in each cover type. We used Program RSW 
(Leban 1999) to conduct the analysis. Results of this anal- 
ysis included a numeric ranking of the different cover 
types according to their relative “preference,” as well as 
a table of pair-wise comparisons (Mests) indicating the 
degree to which preference differed between types. 

We used paired f-tests to determine if the distribution 
of foraging or roosting locations differed from random 
locations relative to elevation, distance to the nearest 
stream, or distance to the nearest open area (dear-cuts/ 
non-forest in Table 1). For these analyses, we computed 


December 2005 


Spotted Owl Habitat Use 


369 



Figure 2. Observation periods of 20 radio-marked northern Spotted Owls observed on the Olympic Peninsula, 
Washington, 1987-89. Vertical lines indicate intervals used for calculation of seasonal ranges. 


mean expected values from a random sample of 200 lo- 
cations in forest areas in the 100% MCP home range of 
each owl. We used digital stream layers and elevation lay- 
ers in GIS to compute elevation and distance to the near- 
est stream for each owl location and each random loca- 
tion. 

Based on a preliminary analysis of our data, the Wash- 
ington State Forest Practices Board (1996) adopted land 
management guidelines in which they stipulated that 
land managers should maintain a minimum of 5863 acres 
(2372 ha) of “suitable habitat” within a 4.3-km radius 
around Spotted Owl site centers (known or suspected 
nest areas) on the Olympic Peninsula. To evaluate the 
amount of protection afforded by these guidelines, we 
examined the proportion of each cumulative owl home 
range that fell within a 4.3-km radius of the nest area or 
main roost area of each owl, and we compared median 
and mean areas of “suitable habitat” in cumulative owl 
ranges with the target in the Forest Practices Rules. All 
means are expressed as x ± 1 SE. 

Results 

Sample Size and Tracking Periods. We moni- 
tored 22 owls in 12 territories, including 10 resi- 
dent pairs, one territory where we marked one 
member of a resident pair, and one territory where 
we marked an adult female that did not appear to 
have a mate. We did not use the data from the 
unpaired adult female because she did not exhibit 
site fidelity. We also did not use data from one fe- 
male that died shortly after she was radio-marked. 


Of the 11 pairs in which one or both members 
were radio-marked, three nested during the study, 
including one pair in 1987 and two pairs in 1988. 

On average, we tracked individual owls for 438 
± 34 d (range = 166-711 d; Fig. 2). Total reloca- 
tions per owl, not counting incubation locations, 
averaged 366 ± 35 (range = 126-685). Of 3262 
roost locations, we estimated 2360 (72%) by tri- 
angulation and located 902 (28%) by homing in 
on transmitters to locate owls visually in their roost 
trees. 

Annual Ranges. Median estimates of annual 
ranges of individual owls were 1147 ha (75% FK), 
2406 ha (95% FK), and 2290 ha (100% MCP; Table 
2) . Most mean estimates of ranges were larger than 
medians because means were skewed by a few in- 
dividuals with large ranges (Table 2). All owls with 
annual MCP or FK ranges >5000 ha were individ- 
uals that expanded their ranges substantially dur- 
ing fall and winter. Annual ranges were smaller for 
females than for males, with the exception of the 
75% FK estimate (Table 2). In six cases in which 
we monitored owls for two yr, the sequential an- 
nual ranges overlapped by 70 ± 6% based on the 
75% FK (range = 18-100%), 73 ± 5% based on 
the 95% FK (range = 27-100%), and 73 ± 4% 
based on the MCP (range = 38-100%). 



370 


Forsman et al. 


VoL. 39, No. 4 


Table 2. Estimates of annual home-range areas of individual northern Spotted Owls on the Olympic Peninsula, 
Washington, 1987-89. Estimates include the 100% minimum convex polygon (MCP) and the 75% and 95% isopleths 
of the Fixed Kernel (FK) . 


Number of Days and Number of Locations 
IN Sample Period*^ 


Home Range Estimates (ha) 


AND Year* 

Days 

Locations 

Locations 

75% FK 

95% FK 

100% MCP 

BPF87 

410 

160 

191 

771 

1696 

1779 

BPF88 

116 

66 

70 

2209 

4196 

3295 

BPM87 

386 

147 

180 

1935 

4865 

6122 

BPM88 

325 

144 

145 

3106 

8469 

8351 

BRF87 

302 

51 

136 

415 

1189 

1402 

BRF88 

142 

39 

58 

804 

1720 

1151 

BRM87 

373 

52 

172 

532 

1172 

1367 

BRM88 

123 

26 

37 

530 

1045 

683 

ECF87 

354 

72 

129 

7927 

15 212 

10 704 

ECF88 

276 

100 

128 

1404 

4207 

6668 

ECM87 

370 

68 

159 

1108 

2104 

1917 

ECM88 

65 

55 

62 

982 

1876 

1294 

FBF87 

166 

73 

79 

1586 

3483 

3086 

FBM87 

206 

99 

106 

1202 

2411 

2230 

HRM88 

366 

201 

186 

2465 

6924 

7954 

LBF88 

345 

137 

143 

2195 

4931 

4950 

LBM88 

391 

168 

184 

1513 

3232 

3288 

LCF88 

246 

65 

109 

967 

2235 

1915 

LCM87 

354 

45 

140 

678 

1504 

2000 

LCM88 

94 

23 

42 

509 

1074 

894 

NRF87 

387 

133 

174 

734 

1786 

2350 

NRF88 

95 

58 

58 

2597 

5795 

4537 

NRF88 

95 

58 

58 

2597 

5795 

4537 

NRM87 

387 

165 

210 

1186 

3084 

4284 

NRM88 

274 

153 

157 

5003 

11558 

11252 

RFF88 

348 

208 

190 

980 

2092 

2235 

RFF89 

114 

50 

44 

554 

1294 

975 

RFM88 

309 

152 

144 

2516 

7059 

6704 

SPF87 

396 

154 

169 

468 

1115 

1323 

SPF88 

86 

48 

55 

1228 

2402 

1406 

SPM87 

396 

154 

167 

568 

1583 

1861 

SPM88 

247 

151 

142 

1072 

2533 

3593 

SRM87 

314 

34 

92 

2817 

5693 

3879 

Mean 




1642 

3736 

3608 

Median 




1147 

2406 

2290 

Mean 9 ^ 




1656 

3557 

3185 

Mean 




1631 

3893 

3981 

Median 




967 

2235 

1915 

Median d'* 




1186 

2533 

3288 


First two letters indicate owl name, third letter indicates sex of owl, and numbers indicate year of estimate. 
Total locations for NRF87 and SPF87 also included 63 and 76 incubation locations, respectively. 

TV = 1 6 owl years. 

TV = 17 owl years. 


December 2005 


Spotted Owl Habitat Use 


371 


Table 3. Estimates of the cumulative home-range areas of northern Spotted Owls on the Olympic Peninsula, Wash- 
ington, 1987-89. Estimates include the 100% minimum convex polygon (MCP) and the 75% and 95% isopleths of 
the Fixed Kernal (FK) estimator. 


Number of Days and Number of Locations 
IN Sample Period 


Owl Code 
Name, Sex 

Days 

Roost 

Locations 

Forage 

Locations 

Home Range Estimates (ha) 

75% FK 95% FK 100% MCP 

BPF 

526 

226 

261 

2216 

4303 

3527 

BPM 

711 

291 

325 

3235 

8521 

8715 

BRF 

444 

90 

194 

804 

1746 

1562 

BRM 

496 

78 

209 

645 

1281 

1372 

ECF 

630 

172 

257 

7927 

15212 

10916 

ECM 

435 

134 

221 

1215 

2166 

1932 

LCM 

449 

68 

182 

845 

1636 

2026 

NRF^ 

482 

191 

232 

2580 

5995 

4852 

NRM 

661 

318 

367 

5003 

11561 

11252 

RFF 

462 

258 

234 

985 

2164 

2298 

SPF^ 

482 

202 

224 

1230 

2436 

1831 

SPM 

643 

305 

309 

1090 

2643 

3716 

Mean^ 




2315 

4972 

4500 

Median*" 




1222 

2539 

2912 

Mean 




2624 

5309 

4164 

Mean 




2006 

4635 

4836 

Median 9'^ 




1173 

3384 

2912 

Median 




1152 

2404 

2870 


^ Total locations for NRF and SPF also included 62 and 76 incubation locations, respectively. 
bN = 12. 

= 6. 


Overlap of annual ranges of nine owls of the 
same sex that occupied adjacent territories aver- 
aged 5 ± 2% for the 75% FK (range = 0-25%), 
21 ± 3% for the 95% FK (range = 3-58%), and 
26 ± 4% for the MCP (range = 0-58%). These 
estimates probably did not reflect total overlap 
with adjacent residents because there were adja- 
cent pairs that we did not have radio-marked and 
because tracking periods for individual owls were 
not always exactly the same. However, even with 
incomplete data on some individuals and no data 
on the pairs that were not radio-marked, it was 
clear that home ranges of neighbors overlapped 
considerably, particularly during winter. In one 
case, a male from one territory (BPM) was found 
on several occasions during winter, roosting in the 
traditional nest area of an adjacent male (HRM) . 

Cumulative Ranges. Median estimates of cumu- 
lative ranges of individual owls monitored in two 
sequential years were 1222 ha (75% FK), 2539 ha 
(95% FK), and 2912 ha (MCP; Table 3). Cumula- 
tive ranges of females averaged larger than cumu- 


lative ranges of males for all comparisons except 
the mean MCP (Table 3). 

Seasonal Ranges. Ranges of individual owls 
based on the 95% FK averaged 3360 ± 572 ha dur- 
ing the breeding season (range = 883—10 205 ha, 
median = 2052 ha, = 21) and 3175 ± 572 ha 
during the nonbreeding season (range = 611-12 
352 ha, median — 2168 ha, N = 29). There was no 
consistent pattern of larger ranges in one season 
or the other. Median estimates of seasonal ranges 
were smaller than means because means were pos- 
itively skewed by a few individuals with large rang- 
es. Overlap of nonbreeding season ranges on 
breeding season ranges averaged 65 ± 4.5% 
(range = 8-98%, N= 36), and overlap of breeding 
season ranges on nonbreeding ranges averaged 63 
± 4.9% (range = 1-100%, N = 36). Overlap of 
breeding season ranges of two owls tracked in two 
different breeding seasons averaged 74 ± 8.9% 
(range = 58-91%). Overlap of nonbreeding rang- 
es of nine owls tracked in two different nonbreed- 


372 


Forsman et al. 


VoL. 39, No. 4 


Table 4. Results of compositional analysis of habitat use for foraging by northern Spotted Owls on the Olympic 
Peninsula, Washington, 1987-89. Rank scores indicate relative preference of cover types from highest (6) to lowest 
(0) . Results of pairwise t-tests indicate the relative preference of cover types. A positive lvalue indicates that the row 
cover type ranked higher than the column cover type and a negative t-value indicates that the row cover type ranked 
lower than the column cover type. A significant P-value suggests that confidence in the direction of the relationship 
was high. 


Cover Type^ 


Old 

Forest 

Mature 

Forest 

Mixed- 

Young 

Forest 

Young 

Forest 

Pole- 

sapling 

Hard- 

wood/ 

Riparian 

Clear- 

cut/Non- 

forest 

Rank 

Old Forest 

t 


3.127 

4.459 

4.637 

8.427 

4.443 

8.103 

6 


P 


0.006 

<0.001 

<0.001 

<0.001 

<0.001 

<0.001 


Mature Forest 

t 

-3.127 


-1.454 

0.774 

2.429 

0.391 

4.183 

4 


P 

0.006 


0.162 

0.448 

0.025 

0.700 

0.001 


Mixed-young Forest 

t 

-4.459 

1.454 


2.756 

6.676 

2.743 

6.528 

5 


P 

<0.001 

0.162 


0.013 

<0.001 

0.013 

<0.001 


Young Forest 

t 

-4.637 

-0.774 

-2.756 


1.561 

-0.471 

2.631 

2 


P 

<0.001 

0.448 

0.013 


0.135 

0.643 

0.017 


Pole-sapling 

t 

-8.427 

-2.429 

-6.676 

-1.561 


-2.826 

2.370 

1 


P 

<0.001 

0.025 

<0.001 

0.135 


0.011 

0.029 


Hardwood/Riparian 

t 

-4.443 

-0.391 

-2.743 

0.471 

2.826 


3.860 

3 


P 

<0.001 

0.700 

0.013 

0.643 

0.011 


0.001 


Clear-cut/Non-forest 

t 

-8.102 

-4.183 

-6.528 

-2.631 

-2.370 

-3.860 


0 


P 

<0.001 

0.001 

<0.001 

0.017 

0.029 

0.001 




ing seasons averaged 59 ± 6.3% (range = 10- 

100 %). 

During the breeding season, movements of owls 
were typically centered on the nest tree or, in the 
case of nonnesting pairs, a regularly-used roost 
area. Winter ranges typically included part of the 
breeding-season range plus areas peripheral to the 
breeding-season range. However, a few individuals 
spent little time in their breeding-season ranges 
during the winter season. The most dramatic ex- 
ample was the Elk Creek Female (ECF) . After nest- 
ing and producing a juvenile in 1987, she left the 
nest area in August and spent most of the fall and 
winter in an area 5-15 km away from the nest area 
before eventually returning to the nest area in 
June of 1988. The Neilton Ridge Male (NRM) also 
had a very large nonbreeding range in 1988-89, 
but in his case, the nonbreeding range overlapped 
most of the breeding season range. 

Ranges of Pairs. There were 14 cases where we 
monitored annual ranges of paired owls in the 
same year. The annual ranges of these pairs (union 
of annual ranges of male and female) averaged 
2397 ± 558 ha for the 75% FK (median = 1570 
ha), 5449 ± 1111 ha for the 95% FK (median = 
4081 ha), and 5414 ± 895 ha for the MCP (median 
= 5032 ha). Overlap of annual ranges of paired 


owls averaged 70 ± 5% based on the 75% FK 
(range = 14-100%), 69 ± 5% based on the 95%FK 
(range = 14—100%), and 64 ± 5% based on the 
MCP (range = 14-100%). Estimates of mean over- 
lap of annual ranges were similar, regardless of 
which sex was used as the frame of reference, so 
we based the above averages on all possible com- 
binations of overlap. 

Cumulative ranges of five pairs that were moni- 
tored in both years averaged 3945 ± 1282 ha for 
the 75% EK (median = 4053 ha), 8278 ± 2550 ha 
for the 95% EK (median = 9329 ha), and 7488 ± 
1951 ha for the MCP (median = 9195 ha). Overlap 
of cumulative 95% FK ranges of paired individuals 
averaged 68 ± 14% for males on females and 72 
± 12% for females on males. 

Habitat Selection. Use of cover types for forag- 
ing and roosting was nonrandom. Old Forest was 
the most preferred type for foraging, followed by 
Mixed-young Forest, Mature Forest, Hardwood/Ri- 
parian Forest, Young Forest, Pole-sapling, and 
Clear-cut/Non-forest (Table 4). Pairwise compari- 
sons of rank indicated that Old Forests were con- 
sistently preferred over all other cover types (Table 
4). Although Mixed-young Forest ranked higher 
than Mature Forest, pairwise comparisons of rank 
indicated little difference between the two types 


December 2005 


Spotted Owl Habitat Use 


373 


Table 5. Results of compositional analysis of habitat use for roosting by northern Spotted Owls on the Olympic 
Peninsula, Washington, 1987-89. Rank scores indicate relative preference of cover types from highest (6) to lowest 
(0) . Results of pairwise t-tests comparisons indicate the relative preference of cover types. A positive lvalue indicates 
that the row cover type ranked higher than the column cover type and a negative f-value indicates that the row cover 
type ranked lower than the column cover type. A significant P-value suggests that confidence in the direction of the 
relationship was high. 


Cover Type^* 


Old 

Forest 

Mature 

Forest 

Mixed- 

Young 

Forest 

Young 

Forest 

Pole- 

sapling 

Hard- 

wood/ 

Riparian 

Clear- 

cut/Non- 

forest 

Rank 

Old Forest 

t 


2.605 

3.326 

4.823 

8.079 

4.752 

16.554 

6 


P 


0.017 

0.004 

<0.001 

<0.001 

<0.001 

<0.001 


Mature Forest 

t 

-2.605 


0.121 

1.141 

5.124 

1.361 

10.527 

5 


P 

0.017 


0.905 

0.268 

<0.001 

0.189 

<0.001 


Mixed-young Forest 

t 

-3.326 

-0.121 


1.163 

5.195 

1.711 

9.010 

4 


P 

0.004 

0.905 


0.259 

<0.001 

0.103 

<0.001 


Young Forest 

t 

-4.823 

-1.141 

-1.163 


3.541 

0.447 

7.540 

3 


P 

<0.001 

0.268 

0.259 


0.002 

0.660 

<0.001 


Pole-sapling 

t 

-8.079 

-5.124 

-5.195 

-3.541 


-3.271 

2.543 

1 


P 

<0.001 

<0.001 

<0.001 

0.002 


0.004 

0.020 


Hardwood/Riparian 

t 

-4.752 

-1.361 

-1.711 

-0.447 

3.271 


6.477 

2 


P 

<0.001 

0.189 

0.103 

0.660 

0.004 


<0.001 


Clear-cut/Non-forest 

t 

-16.554 

-10.527 

-9.010 

-7.540 

-2.543 

-6.477 


0 


P 

<0.001 

<0.001 

<0.001 

<0.001 

0.020 

<0.001 




(Table 4). Similarly, Young Forest ranked lower 
than Mature and Hardwood Forest, but pairwise 
comparisons indicated that these differences were 
weak (Table 4). Pole-sapling stands ranked lower 
than all other types except Clear-cuts, but the pair- 
wise comparisons with other types indicated that 
preference for Pole-sapling was not greatly differ- 
ent from Young Forest (Table 4). Large P-values 
for all pairwise comparisons of Clear-cuts relative 
to other cover types indicated that Clear-cuts were 
the least preferred cover type for foraging. In fact, 
out of 3822 foraging locations where cover type 
could be determined, only 57 (1.5%) occurred in 
Clear-cuts or Non-forest areas, and we suspected 
that some of these cases were due to telemetry or 
mapping error. 

Use of cover types for roosting indicated that 
Old Forests were preferred over all other cover 
types (Table 5) . Mature Forest ranked higher than 
Mixed-young, Young Forest, and Hardwood/Ripar- 
ian Forest, but pairwise comparisons of these types 
indicated that differences among them were weak 
(Table 5). Pole-sapling, Clear-cuts, and Non-forest 
areas were rarely used for roosting. Of 902 roosts 
located visually, none were located in Clear-cuts or 
Non-forest. Of 2275 roosts located by triangulation 
alone, and for which cover type was determined, 


eight were in Clear-cuts or Non-forest types; we sus- 
pected these were due to triangulation or mapping 
error. 

Habitat Use Relative to Forest Edges, Streams, 
and Elevation. On average, foraging locations and 
roost locations were closer to openings (233 ± 24 
m, and 271.9 ± 33.0 m, respectively) than were 
random locations (304 ± 34 m; borage ~ — 4.10, P 
= 0.001, ^roost = -2.04, P = 0.055; - 20 owls). 

However, the number of locations within 100 m of 
an edge was similar between random locations and 
foraging locations (28.4% vs. 33.5%) and random 
locations and roost locations (28.4% vs. 29.9%), so 
we concluded that there was little evidence that 
owls either preferred or avoided forest edges for 
roosting or foraging. 

Mean elevations at foraging locations (315 ± 29 
m) and roosting locations (322 ± 31 m) were 
slightly lower than elevations at random locations 
(354 ± 36 m; = -3.63, P = 0.002, W ^ 
— 3.09, P = 0.006, N — 20 owls). Mean distance to 
the nearest stream was similar for foraging (98 ± 
14 m), roosting (112 ± 19 m), and random loca- 
tions (94 ± 10 m; ^forage — 0.73, P = 0.475, ~ 

1.75, P = 0.097, = 20 owls). 

Landscape Composition and Home Range Size. 
Size of annual ranges was negatively correlated 


374 


Forsman et al. 


VoL. 39, No. 4 


with the percent cover of older forest (cover types: 
Old and Mature forest) in the cumulative MCP 
range, regardless of whether the estimator was the 
75% FK (rgi = -0.53, P = 0.002), 95% FK (r^i = 
-0.59, P < 0.001), or MCP = -0.67, P < 
0.001). Size of annual ranges was also negatively 
correlated with the amount of older forest in a 4.3 
km circle centered on the central place (75% FK 
r3i - -0.34, P = 0.058; 95% FK - -0.40, P - 
0.028; MCP = -0.46, P - 0.009). 

Overlap of Management Circles with Home 
Ranges. On average, a 4.3-km radius circle cen- 
tered on the nest site or center of activity included 
94 ± 2% of the annual 75% FK home range, 86 ± 
4% of the annual 95% FK home range, and 83 ± 
4% of the annual MCP range. For 12 owls tracked 
in both years, average overlap of the 4.3-km radius 
circle on the cumulative range was 99 ± 13% for 
the 75% FK range, 79 ± 7% for the 95% FKrange, 
and 76 ± 7% for the MCP range. The counter- 
intuitive result in which overlap of the 4.3-km cir- 
cle was lower on the 75% FK annual range than 
on the 75% FK cumulative range occurred because 
the estimates were based on different individuals. 
If we defined “suitable habitat” as the cover types 
that had the top three preference rankings based 
on compositional analysis (cover types = Old, Ma- 
ture, and Mixed-young Forest), then the mean 
amount of suitable habitat within a 4.3-km radius 
circle was 3105 ± 236 ha. 

Discussion 

Home Range Attributes. The large ranges ob- 
served in our study suggest that biomass of suitable 
prey for Spotted Owls is lower on the Olympic Pen- 
insula than in western Oregon and northwestern 
California, where home ranges tend to be smaller 
(Forsman et al. 1984, Carey et al. 1990, 1992, Zabel 
et al. 1995, Bingham and Noon 1997, Glenn et al. 
2004) . We did not have data on total prey biomass 
in our study area, but Carey et al. (1992) found 
that flying squirrels, which are the primary prey of 
Spotted Owls on the Olympic Peninsula, were rel- 
atively uncommon on the peninsula compared to 
western Oregon. 

As in our study, Carey et al. (1990) and Glenn 
et al. (2004) found that home range size of north- 
ern Spotted Owls was inversely related to the 
amount of old forest in the home range. This sug- 
gests that Spotted Owls respond to decreasing 
amounts of their preferred habitat by increasing 
the size of their ranges to encompass more old for- 


est. However, Zabel et al. (1995) found no corre- 
lation between home-range size of Spotted Owls 
and the proportion of the range covered by large 
trees. Instead, they found that home-range size was 
positively correlated with the proportion of flying 
squirrels in the diet and negatively correlated with 
the proportion of woodrats (Neotoma spp.) in the 
diet. In our study area, the diet was dominated by 
flying squirrels (Forsman et al. 2001), which tend 
to be most abundant in old forests (Carey et al. 
1992, Waters and Zabel 1995). This could explain 
why home ranges in our study area became larger 
as the amount of old forest declined. However, for 
a central-place forager like the Spotted Owl, the 
ability to increase the size of the home range and 
still function as a part of the resident breeding 
population is probably limited by energetic and so- 
cial constraints (Carey et al. 1992). 

In our study, annual home ranges of paired owls 
typically overlapped by 50-80%. Similar estimates 
were obtained in a number of other studies (Fors- 
man et al. 1984, Carey et al. 1990, Glenn et al. 
2004). Our estimates of mean overlap of annual 
ranges of owls on adjacent territories were higher 
than values reported by Forsman et al. (1984:23; 
MCP overlap = 12%) and Glenn et al. (2004:41; 
95% FK overlap = 14.9 ± 4.3% and 6.7 ± 2.2% 
on two different study areas) . 

Habitat Selection. Our study, and most other 
studies in which telemetry methods have been 
used to examine habitat selection by northern 
Spotted Owls, indicated that, given a choice, most 
individuals selectively used older forests for forag- 
ing and roosting and that younger stands generally 
provided lower quality habitat (e.g., Forsman et al. 
1984, Call 1989, Carey et al. 1990, 1992, Solis and 
Gutierrez 1990, Gutierrez et al. 1995). However, 
there have been two radiotelemetry studies of 
northern Spotted Owls in landscapes dominated 
by young forest, where patterns of habitat selection 
were less clear. Glenn et al. (2004) examined hab- 
itat selection by Spotted Owls in young forests in 
northwest Oregon and did not find strong selec- 
tion for any cover type. In a landscape where old 
forest comprised less than 10% of the available cov- 
er, Irwin et al. (2000) found that northern Spotted 
Owls infrequently used stands <25 yr of age and 
foraged primarily in mid-age stands (25—79 yr old) 
or in remnant patches of old forest. However, Ir- 
win et al. (2000) did not conduct a landscape-level 
analysis of use-versus-availability with their data, so 


December 2005 


Spotted Owl Habitat Use 


375 


we could not determine if use of different cover 
types differed from availability. 

California Spotted Owls (5. o. occidentalis) in the 
Sierra Nevada Mountains tended to forage in for- 
ests with ^40% canopy cover, but did not show a 
strong preference relative to tree age or tree size 
(Zabel et al. 1992). However, at two of the three 
study areas described by Zabel et al. (1992), the 
majority of foraging and roosting locations were in 
stands dominated by large (>53 cm DBH) trees. 

Of the 5-6 species of small mammals that com- 
prise the primary diet of Spotted Owls, several ap- 
pear to be most abundant in older forests. For ex- 
ample, there are a number of studies that suggest 
that red tree voles {Arborimus longicaudus) and red- 
backed voles {Clethrionomys californicus) are most 
abundant in older forests (Corn and Bury 1986, 
Aubry et al. 1991, Rosenberg et al. 1994). While 
not all studies of northern flying squirrels have 
found significantly higher numbers in old forests, 
the trend in most studies was toward higher num- 
bers in old forests (Carey et al. 1992, Rosenberg 
and Anthony 1992, Waters and Zabel 1995, Lehm- 
kuhl et al. (in press). Therefore, an obvious hy- 
pothesis is that differences in abundance of pre- 
ferred prey cause northern Spotted Owls to select 
for older forests (Forsman et al. 1984, Carey et al. 
1992). Ward et al. (1998) posed a similar hypoth- 
esis to explain high use of forest edges by Spotted 
Owls in northwestern California, where the diet 
was dominated by dusky-footed woodrats {N. fusci- 
pes), which were most abundant in brushy open- 
ings adjacent to forests. In contrast, in areas where 
they feed mainly on flying squirrels. Spotted Owls 
either avoid non-forest edges or use them in pro- 
portion to availability (Zabel et al. 1995, Glenn et 
al. 2004, this study). 

Streams and Elevation. Although Glenn et al. 
(2004) found evidence that Spotted Owls foraged 
selectively in riparian vegetation, we found no ev- 
idence that foraging or roosting locations were 
closer to streams than were random locations. We 
concluded that there was no evidence from our 
data that owls were either selecting or avoiding ri- 
parian areas. Although Spotted Owls in our study 
foraged at lower elevations than expected, the 
mean difference between observed and expected 
foraging locations was only 39 m. We were not con- 
vinced that this relatively small difference was bi- 
ologically meaningful. 

Management Implications, Based on the results 
of our study, we agree with Forsman et al. (1984), 


Thomas et al. (1990), and Carey et al. (1992) that 
management for northern Spotted Owls in western 
Washington and Oregon should focus on the re- 
tention of old forests. Although Franklin et al. 
(2000) and Olson et al. (2004) found that north- 
ern Spotted Owls may have higher reproductive 
output in landscapes that include a mixture of old 
forest and edges with other forest types, those stud- 
ies were conducted in areas where woodrats were 
a primary prey, and the results may not apply to 
areas like the Olympic Peninsula, where flying 
squirrels are the primary prey. 

Bingham and Noon (1997, 1998) suggested that 
the U.S. Fish and Wildlife Service should focus on 
the most heavily-used portion of the home range, 
or “core area,” as the frame of reference for as- 
sessment of “take” of Spotted Owls. If this ap- 
proach is used on the Olympic Peninsula, then we 
believe it would be reasonable to use our estimates 
of the 75% isopleth of the FK annual range as the 
criteria for estimates of core areas, although other 
methods have been proposed (Bingham and Noon 
1997). We agree with Bingham and Noon (1997) 
that it makes sense to use repeatable measures of 
home range areas as the frame of reference for 
assessments of “take,” but this should not be mis- 
construed as a recommendation to manage Spot- 
ted Owls based only on core areas. If the objective 
is to provide Spotted Owls with enough habitat to 
survive and reproduce on a site, then we agree with 
Buchanan et al. (1998) that management should 
be based on amounts of habitat within the entire 
home-range areas of radio-marked owls, not just 
core areas. 

Our estimates of the median and mean amounts 
of “suitable habitat” within cumulative MCP rang- 
es of Spotted Owls (1824 ha and 2253 ± 286 ha) 
are similar to or slightly lower than the manage- 
ment target adopted by the Washington State For- 
est Practices Board (1996) for management 
around Spotted Owl nest sites (2373 ha of suitable 
habitat within a 4.3-km radius). We found that a 
4.3-km radius circle centered on the nest site en- 
compassed about 83-87% of the mean cumulative 
home range used by individual Spotted Owls on 
the peninsula. Based on these results, we see no 
reason to suggest changes to the 1996 Forest Prac- 
tices Rules (Washington State Forest Practices 
Board 1996). However, it remains to be seen if 
Spotted Owls will persist in areas where old and 
mature forests are gradually replaced with less-pre- 
ferred types that are also classified as “suitable.” 


376 


Forsman et al. 


VoL. 39, No. 4 


Acknowledgments 

For help with data collection, we thank Duane Aubu- 
chon, Bruce Casler, Sue Grayson, Martha Jensen, Pete 
Loschl, Rich Lowell, Paul Radley, Doreen Schmidt, Margy 
Taylor, and Joe Zisa. Housing and administrative support 
was provided by the Olympic National Forest. Reviews of 
the manuscript by Rocky Gutierrez, Martin Raphael, Dan 
Varland, and Cindy Zabel were extremely helpful. This 
study was funded by the U.S. Department of Agriculture 
Forest Service, Pacific Northwest Research Station, Port- 
land, OR U.S.A. 

Literature Cited 

Aebischer, N.J., V. Marcstrom, R.E. Kenward, and M. 
Karlbom. 1993. Survival and habitat utilization: a case 
for compositional analysis. Pages 343-353 mJ.D. Le- 
breton and P.M. North [Eds.], Marked individuals in 
the study of bird population. Birkhauser Verlag, Basel, 
Switzerland. 

Aubry, K.B, MJ. Crites, and S.D. West. 1991. Regional 
patterns of small mammal abundance and community 
composition in Oregon and Washington. Pages 285- 
294 in L.F. Ruggerio, K.B. Aubry, A.B. Carey, and 
M.H. Huff [Tech. Coords.], Wildlife and vegetation 
of unmanaged Douglas-fir forests. Gen. Tech. Rep. 
PNW-GTR-285. USDA, Forest Service, Pacific North- 
west Research Station, Portland, OR U.S.A. 

Bingham, B.B. and B.R. Noon. 1997. Mitigation of hab- 
itat “take”: application to habitat conservation plan- 
ning. Conserv. Biol. 11:127-139. 

AND . 1998. The use of core areas in com- 
prehensive mitigation strategies. Conserv. Biol. 12:241- 
243. 

Buchanan, J.B., FJ. Frederickson, and D.E. Seaman. 
1998. Mitigation of habitat “take” and the core area 
concept. Conserv. Biol. 12:238—240. 

Call, D.R. 1989. Home range and habitat use by Cali- 
fornia Spotted Owls in the central Sierra Nevada. M.S. 
thesis, Humboldt State University, Areata, CA U.S. A. 
Carey, A.B., S.P. Horton, and B.L. Biswell. 1992. North- 
ern Spotted Owls: influence of prey base and land- 
scape character. Ecol. Monogr. 62:223-250. 

AND K.C. Peeler. 1995. Spotted Owls: resource 

and space use in mosaic landscapes. J. Raptor Res. 29: 
223-229. 

, J.A. Reid, and S.P. Horton. 1990. Spotted Owl 

home range and habitat use in southern Oregon 
Coast Ranges. /. Wildl. Manage. 54:11-17. 

Corn, P.S. and R.B. Bury. 1986. Habitat use and terres- 
trial activity by red tree voles {Arborimus longicaudus) 
in Oregon./. Mammalogy 67:404—406. 

Forsman, E.D. 1983. Methods and materials for locating 
and studying Spotted Owls. Gen. Tech. Rep. PNW- 
162. USDA, Forest Service, Pacific Northwest Re- 
search Station, Portland, OR U.S.A. 

, E.C. Meslow, and H.M. Wight. 1984. Distribu- 
tion and biology of the Spotted Owl in Oregon. Wildl. 
Monogr. 87:1—64. 


, LA. Otto, S.G. Sovern, M. Taylor, D.W. Hays, 

H. Allen, S.L. Roberts, and D.E. Seaman. 2001. Spa- 
tial and temporal variation in diets of Spotted Owls 
in Washington./. Raptor Res. 35:141-150. 

Franklin, A.B., D.R. Anderson, RJ. Gutierrez, and K.P. 
Burnham. 2000. Climate, habitat quality, and fitness 
in Northern Spotted Owl populations in northwest- 
ern California. Ecol. Monogr. 70:539-590. 

Glenn, E.M., M.C. Hansen, and R.G. Anthony. 2004. 
Spotted Owl home range and habitat use in young 
forests of western Oregon. /. Wildl. Manage. 68:33-50. 

Guetterman, J.H., J.A. Burns, J.A. Reid, R.B. Horn, and 
C.C. Foster. 1991. Radiotelemetry methods for study- 
ing Spotted Owls in the Pacific Northwest. Gen. Tech. 
Rep. PNW-272. USDA, Forest Service, Pacific North- 
west Research Station, Portland, OR U.S.A. 

Gutierrez, R.J., A.B. Franklin, and W.S. LaHaye. 1995 
Spotted Owl. The Birds of North America No. 179. 
The Birds of North America, Inc. Philadelphia, PA 
U.S.A. 

Hamer, T.E. 1988. Home range size of the Northern 
Barred Owl and Northern Spotted Owl in western 
Washington. M.S. thesis. Western Washington Univer- 
sity. Seattle, WA U.S.A. 

Havne, D.W. 1949. Calculation of size of home range. /. 
Mammalogy 30:1-18. 

Henderson, J.A., D.H. Peter, R.D. Lesher, and D.C 
Shaw. 1986. Forested plant associations of the Olym- 
pic National Forest. Ecological Technical Paper R6- 
ECOL-TP 001-88. USDA, Forest Service, Portland, OR 
U.S.A. 

Irwin, L.L., D.R Rock, and G.P. Miller. 2000. Stand 
structures used by northern Spotted Owls in managed 
forests./. Raptor Res. 34:175-186. 

Kie,J.G., J.A. Baldwin, and C.J. Evans. 1996. CALHOME- 
A program for estimating animal home ranges. Wildl 
Soc. Bull. 24:342-344. 

Laymon, S.A. 1989. Altitudinal migration movements of 
Spotted Owls in the Sierra Nevada, California. Condor 
91:837-841. 

Leban, F. 1999. Program RSW — Resource selection for 
Windows, Version 1.00 Beta 8.4. University of Idaho, 
Moscow, ID U.S.A. 

Lehmkuhl, J.F., K.D. Kistler, J.S. Begley, and J. Boulan- 
ger. (in press) . Demography of northern flying squir- 
rels informs ecosystem management of western inte- 
rior forests. Ecol, Appl. 

Nams, V.O. and S. Boutin. 1991. What is wrong with er- 
ror polygons? /. Wildl. Manage. 55:172-176. 

Olson, G.S., E.M. Glenn, R.G. Anthony, E.D. Forsman, 
J.A. Reid, PJ. Loschl, and WJ. Ripple. 2004. Model- 
ing demographic performance of northern Spotted 
Owls relative to forest habitat in Oregon. /. Wildl 
Manage. 68:1039—1053. 

Otis, D.L., and G.C. White. 1999. Autocorrelation of lo- 
cation estimates and the analysis of radiotracking 
data.. J. Wildl. Manage. 63:1039-1044. 


December 2005 


Spotted Owl Habitat Use 


377 


Pierce, F.R. 1921. Tornado destroys great forest. Pop. 
Mech. 35:659-662. 

Rosenberg, D.K. and R.G. Anthony. 1992. Characteris- 
tics of northern flying squirrel populations in young 
second- and old-growth forests of western Oregon. 
Can.;. Zool. 70:161-166. 

, K.A. Swindle, and R.G. Anthony. 1994. Habitat 

associations of California red-backed voles in young 
and old-growth forests in western Oregon. Northwest 

Sci. 68:266-272. 

Seaman, D.E., B. Griffith, and R.A. Powell. 1998. KER- 
NELHR: a program for estimating animal home-rang- 
es. WildL Soc. Bull. 26:95-100. 

and R.A. Powell. 1996. An evaluation of the ac- 
curacy of kernel density estimators for home-range 
analysis. Ecology 77:2075-2085. 

Solis, D.M., Jr. and RJ. Gutierrez. 1990. Summer hab- 
itat ecology of northern Spotted Owls in northwestern 
California. Condor 92:739-748. 

SwiHART, R.K. and N.A. Slade. 1985a. Influence of sam- 
pling interval on estimates of home-range size. J. 
WildL Manage. 49:1009-1025. 

and N.A. Slade. 1985b. Testing for independence 

of observations in animal movements. Ecology 66: 
1176-1184. 

Thomas, J.W., E.D. Forsman, J.B. Lint, E.C. Meslow, 
B.R. Noon, and J. Verner. 1990. A conservation strat- 
egy for the northern Spotted Owl: report of the In- 
teragency Scientihc Committee to address the conser- 


vation of the northern Spotted Owl. USDA, Forest 
Service and USDI, Bureau of Land Management, 
Portland, OR U.S.A. 

Ward, J.P., Jr., RJ. Gutierrez, and B.R. Noon. 1998. 
Habitat selection by Northern Spotted Owls: the con- 
sequences of prey selection and distribution. Condor 
100:79-92. 

Washington State Forest Practices Board. 1996. Final 
environmental impact statement on forest practices 
rule proposals for northern Spotted Owl, Marbled 
Murrelet, and western gray squirrel. Washington State 
Forest Practices Board, Olympia, WA U.S.A. 

Waters, J.R. and C.J. Zabel. 1995. Northern flying squir- 
rel densities in fir forests of northeastern California. 
J. WildL Manage. 59:858-866. 

Zabel, C.J., K. McKelvey, and J.P. Ward, Jr. 1995. Influ- 
ence of primary prey on home-range size and habitat- 
use patterns of northern Spotted Owls (Strix occiden- 
talis caurina). Can. J. Zool. 73:433—439. 

, G.N. Steger, K.S. McKelvey, G.P. Eberlein, B.R. 

Noon, and J. Verner. 1992. Home-range size and 
habitat-use patterns of California Spotted Owls in the 
Sierra Nevada. Pages 149-163 in]. Verner, K.S. Mc- 
Kelvey, B.R. Noon, RJ. Gutierrez, G.I. Gould, Jr., and 
T.W. Beck [Eds.], The California Spotted Owl: a tech- 
nical assessment of it current status. USDA, Forest 
Service, Pacific Southwest Research Station, Albany, 
CA U.S.A. 

Received 4 April 2003; accepted 30 August 2005 


J Raptor Res. 39(4):378-385 
© 2005 The Raptor Research Foundation, Inc. 

FIRST-CYCLE MOLTS IN NORTH AMERICAN FALCONIFORMES 

Peter Pyle' 

The Institute for Bird Populations, P.O. Box 1436, Point Reyes Station, CA 94956 U.S.A. 

Abstract. — I examined 1849 specimens of 20 North American Falconiform species to elucidate the 
occurrence and nomenclature of partial first-cycle molts. As reported in the literature, American Kestrel 
{Falco sparverius) and White-tailed Kite (Elanus leucurus) have relatively complete body-feather molts that 
occur during the first fall; in the kite, this molt can also include up to all rectrices and 2-6 secondaries, 
but no primaries — an unusual pattern for such partial molts in first-year birds. Evidence of partial first- 
cycle molts was found in 16 of 18 other species (among Pandion, Haliaeetus, Circus, Accipiter, Asturina, 
Buteo, Aquila, and Falco) for which such molts have not been previously elucidated. Maximum extent of 
body-feather replacement among individuals of these 16 species varied from 5-50%. On the other hand, 
most species showed evidence that this molt could be absent (11-100% of birds remaining in juvenile 
plumage until commencement of the complete or near-complete prebasic molt that occurs during the 
first summer) . I argue that these partial molts are best considered preformative molts (following Howell 
et al. 2003) rather than “first prebasic” molts, as defined by Humphrey and Parkes (1959). Variation 
in the extent and timing of preformative molts may reflect various constraints according to species- 
specihc breeding, migrating, and foraging strategies. The apparent lack of function for this molt suggests 
that ancestral Falconiformes exhibited a more extensive preformative molt, as found in related orders 
of birds, but that this molt has since become vestigial, at least in the larger species. 

Key Words: American Kestrel, Falco sparverius; White-tailed Kite, Elanus leucurus; preformative molt, hawks; 

falcons. 


MUDAS DEL PRIMER CICLO EN FALCONIFORMES NORTEAMERICANOS 

Resumen. — Examine 1849 especimenes de 20 especies norteamericanas de Falconiformes para estable- 
cer la ocurrencia y la nomenclatura de mudas parciales del primer ciclo. Como habia sido informado 
en la literatura, Falco sparverius y Elanus leucurus presentan mudas relativamente completas de las plumas 
corporales que tienen lugar durante el primer otono, aunque en E. leucurus esta muda puede tambien 
incluir algunas o todas las rectrices y 2-6 secundarias pero ninguna primaria. Esto constituye un patron 
poco usual de mudas parciales durante el primer aho de vida. Encontre evidencia de mudas parciales 
en el primer ciclo en 16 de las 18 especies adicionales (en los generos Pandion, Haliaeetus, Circus, Accipiter, 
Asturina, Buteo, Aquila y Falco) para las cuales no se habia determinado previamente la ocurrencia de 
mudas de este tipo. El maximo grado de reemplazo de las plumas del cuerpo en los individuos de esas 
16 especies vario entre el 5% y el 50%. Por otra parte, la mayoria de las especies mostraron evidencia 
de que esta muda podria estar ausente, pues entre el 11% y el 100% de los individuos mantuvieron el 
plumaje juvenil hasta comenzar la muda prebasica completa, o casi completa, que tiene lugar durante 
el primer verano. Sugiero que estas mudas prebasicas deben considerase mudas preformativas (siguien- 
do la terminologia de Howell et al. 2003) en lugar de “primeras mudas prebasicas”, como fueron 
definidas por Humphrey y Parkes (1959). La variacion en la extension y en el momenta de ocurrencia 
de las mudas preformativas podria reflejar distintos limitantes de acuerdo a las estrategias reproductivas, 
de migracion y de forrajeo especificas de cada especie. La falta aparente de funcionalidad de esta muda 
sugiere que los Falconiformes ancestrales exhibian una muda preformativa mas extensiva, como se ha 
documentado en ordenes de aves relacionados, pero que esta muda se ha vuelto vestigial, al menos en 
las especies de mayor tamaho. 

[Traduccion del equipo editorial] 

North American Falconiformes exhibit various most species, Juvenal plumage is reportedly re- 
molt strategies during their first year of life. In tained until the first spring or summer (when a 

year old) , at which point a complete or near-com- 

plete prebasic molt commences, usually with the 

^ Email address: ppyle@birdpop.org shedding of the innermost primaries in Accipitri- 


378 


December 2005 


First-cycle Molts in Falconiformes 


379 


dae or the medial primaries in Falconidae (e.g., 
Miller 1941, Palmer 1988, Forsman 1999, Wheeler 
2003). In American Kestrel (Falco sparverius) and 
certain kites, by contrast, most to all Juvenal body 
feathers are replaced during the first fall, well be- 
fore molt of the primaries commences during the 
following summer (Bent 1937, 1938, Parkes 1955, 
Palmer 1988, Miller and Smallwood 1997). 

For some species of Falconiformes, a limited 
number of body feathers are reported to be re- 
placed during the first fall, winter, or spring, prior 
to the shedding of primaries during the first spring 
or summer. For example, Wheeler (2003) reported 
that in the Red-shouldered Hawk (see Table 1 for 
scientific names), “the first prebasic molt begins 
on the breast, then is noticeable on the back and 
scapulars” before molt of primaries begins, and 
Forsman (1999) reported that in the Northern 
Harrier, “single body-feathers and tail-feathers may 
be replaced from late winter” prior to the com- 
plete molt the following May to October. Similar 
limited molts in fall, winter, or spring have been 
reported for Osprey (Bent 1937), Northern Gos- 
hawk (Bent 1937, Dement’ev and Gladkov 1951, 
Forsman 1999), Gray Hawk (Dickey and van Ros- 
sem 1938, Wheeler 2003), and several species of 
Falconidae (Cramp and Simmons 1980, Forsman 
1999, Wheeler 2003). 

Most authors consider this limited body-feather 
replacement to be the initiation of the prebasic 
molt at a year of age (hereafter, “first complete 
molt”) rather than a separate molt; indeed, Wheel- 
er (2003) even interprets the fall molts in first-year 
American Kestrel and kites as part of the first com- 
plete molt, terminating with the flight feathers the 
following summer. This interpretation appears to 
disregard the additional body molt that occurs in 
one-year-old birds concurrent with flight-feather 
molt in these species (Parkes 1955, Palmer 1988). 
Herremans and Louette (2000), on the other 
hand, document that such molts in certain Old- 
World species of Accipiter are distinct from the first 
complete molt. 

Thus, there remains confusion about both the 
occurrence of partial first-cycle molts in Falconi- 
formes and whether or not body-feather replace- 
ment in the first fall, winter, or spring should be 
considered part of a separate partial molt or as the 
initiation of the first complete molt. To investigate 
the occurrence and extent of partial first-cycle 
molts in Falconiformes, I examined 1227 speci- 
mens of 20 North American species collected dur- 


ing their first year (age == 0-12 mo), prior to ini- 
tiation of primary molt, and 622 specimens 
collected in their second year (12-24 mo old) or 
later. 

Methods 

Specimens of Falconiformes were examined at the Cal- 
ifornia Academy of Sciences (CAS), San Francisco, the 
Museum of Vertebrate Zoology (MVZ), Berkeley, and the 
National Museum of Natural History (USNM), Washing- 
ton, DC. Specimens were collected throughout North, 
Central, and South America, but all specimens repre- 
sented species or subspecies exhibiting boreal breeding 
cycles. Data are presented for 20 species with A s 16 
first-year individuals examined (Table 1). 

Specimens examined included birds collected during 
their first year of life, prior to the shedding of primaries 
during the first complete molt, and birds determined to 
be in their second year of life. Age of first-year birds was 
determined by plumage features (Palmer 1988, Wheeler 
2003), the presence of indicative fault bars (Hammers- 
trom 1967), tapered and relatively worn outer primaries 
and rectrices, and absence of wear and color patterns 
among flight feathers indicating previous molts. Birds in 
their second year were aged by the presence of predefi- 
nitive plumage in some species or the retention of Juve- 
nal feathers during the first “complete” molt, particular- 
ly among the lesser coverts, on the rump, or within the 
secondaries (Wheeler 2003). 

Each first-year specimen was examined carefully for re- 
placed body feathers in patterns indicating molt. Body 
molt in Falconiformes typically begins on the head and 
throat and proceeds caudally (Palmer 1988, Wheeler 
2003). Thus, feathers showing wear patterns suggesting 
replacement in a caudal direction were assumed to have 
resulted from molt rather than adventitious replacement 
(e.g., after accidental loss). Replaced feathers were often 
intermediate in color or patterning between those of 
first-year and second-year birds (Fig. 1), facilitating their 
identification. The proportion of newly-replaced body 
feathers, to the nearest 5%, was estimated for each spec- 
imen. Those showing <2.5% replacement were scored as 
0%, further ensuring that birds with adventitiously re- 
placed feathers were not included in the sample evincing 
molt. All primaries, secondaries, and rectrices were also 
examined for evidence of symmetrical replacement in- 
dicating molt. Secondaries were numbered proximally 
from the outermost (si) to the innermost (si 3 in most 
species) feather. 

Second-year birds were examined for uniformity of 
feather generations, particularly in tracts for which par- 
tial molts were detected in first-year birds. The goal of 
this examination was to assess whether or not feathers 
replaced during the first year, prior to initiation of pri- 
mary molt, had been replayed for a second time during 
the first complete molt. 

Results 

Partial First-year Molts in White-tailed Kite and 
American Kestrel. I examined 22 specimens of 


380 


Pyle 


VoL. 39, No. 4 


Table 1. Proportion of individuals showing evidence of molt and maximum percent of body feathers replaced, 
among specimens of 18 species of North American Falconiformes, during three time periods within the first year 
Values represent sample size of specimens examined, proportion of individuals showing evidence of molt, and max- 
imum proportion of body feathers replaced among individuals sampled. 




September-November 


December-February 


March- 

May 

Species 

N 

Proportion 

Molting 

Maximum 

Percent 

Feathers 

Replaced 

N 

Proportion 

Molting 

Maximum 

Percent 

Feathers 

Replaced 

N 

Proportion 

Molting 

Maximum 

Percent 

Feathers 

Replaced 

Osprey {Pandion 
haliaetus) 

10 

0.40 

10% 

3 

1.00 

20% 

2 

1.00 

35% 

Bald Eagle {Haliaeetus 
leucocephalus) 

8 

0.12 

5% 

5 

0.60 

30% 

3 

1.00 

40% 

Northern Harrier 
{Circus cyaneus) 

35 

0.00 

0% 

18 

0.33 

15% 

16 

0.44 

50% 

Sharp-shinned Hawk 
{Accipiter striatus) 

80 

0.00 

0% 

43 

0.12 

10% 

48 

0.31 

45% 

Cooper’s Hawk 
{Accipiter cooperii) 

59 

0.08 

10% 

26 

0.31 

10% 

14 

0.64 

10% 

Northern Goshawk 
{Accipiter gentilis) 

30 

0.00 

0% 

6 

0.00 

0% 

10 

0.00 

0% 

Gray Hawk 

{Asturina nitida) 

12 

0.00 

0% 

7 

0.14 

5% 

6 

0.33 

10% 

Red-shouldered Hawk 
{Buteo lineatus) 

11 

0.45 

10% 

11 

0.45 

10% 

2 

0.00 

0% 

Broad-winged Hawk 
{Buteo platypterus) 

10 

0.00 

0% 

9 

0.00 

0% 

7 

0.00 

0% 

Swainson’s Hawk 
{Buteo swains oni) 

7 

0.57 

10% 

5 

1.00 

25% 

15 

0.86 

40% 

Red-tailed Hawk 
{Buteo jamaicensis) 

48 

0.21 

5% 

54 

0.26 

10% 

27 

0.37 

20% 

Ferruginous Hawk 
{Buteo regalis) 

23 

0.13 

5% 

10 

0.50 

5% 

5 

0.60 

10% 

Rough-legged Hawk 
{Buteo lagopus) 

16 

0.31 

5% 

10 

0.50 

15% 

4 

1.00 

10% 

Golden Eagle 

{Aquila chrysaetos) 

6 

0.17 

5% 

9 

0.22 

5% 

3 

1.00 

5% 

Merlin {Falco 
columbarius) 

51 

0.04 

5% 

33 

0.33 

15% 

18 

0.50 

50% 

Gyrfalcon {Falco 
rusticolus) 

15 

0.07 

10% 

12 

0.42 

30% 

5 

0.20 

25% 

Peregrine Falcon 
{Falco peregrinus) 

27 

0.19 

10% 

13 

0.46 

25% 

13 

0.85 

20% 

Prairie Falcon 
{Falco mexicanus) 

24 

0.29 

15% 

12 

0.50 

15% 

9 

0.89 

20% 


White-tailed Kite and 183 specimens of American 
Kestrel collected between September and May of 
their first year. A partial to complete body-feather 
molt during the first fall was confirmed for all in- 
dividuals of both of these species. Proportions of 
replaced feathers indicated that, in both species, 
this molt had completed or nearly completed by 


December, with little or no additional molt taking 
place during January-May, until initiation of the 
first complete molt. Examination of six White- 
tailed Kites and 15 American Kestrels undergoing 
their first complete molt and 73 kites and 274 kes- 
trels in definitive plumage (showing uniform body 
plumage) confirmed that another complete body 


December 2005 


First-cycle Molts in Falconiformes 


381 



A B C D 


Figure 1. Pigraent patterns on underpart feathers of 
Cooper’s Hawks. A — Typical Juvenal feather; B — Forma- 
tive feather replaced in September (CAS62145); C — For- 
mative feather replaced in January (CAS33560); D — Typ- 
ical definitive feather. 


molt occurs in these species concurrent with the 
flight-feather molt at a year of age. 

In White-tailed Kite, 13 first-year individuals col- 
lected between January and the initiation of the 
first complete molt (with shedding of the inner- 
most primaries) had replaced a mean 97.7% of 
body feathers. All 13 kites had replaced at least 
some lesser coverts (but no median coverts) and 
the central two to all 12 rectrices {x = 6.9 feath- 
ers) . In addition, five individuals were found that 
had replaced 2-6 secondaries during the partial 
molt, symmetrically on both wings and apparently 
in the same sequence as observed in complete 
molts, as determined from dines in wear. The two 
specimens with the most extensive partial first-year 
molts were collected in California: MVZ24493 (10 
January) and CAS73280 (22 February), each with 
100% of the body feathers, all 12 rectrices, and sl- 
s2, s5, and sll-sl3 replaced. No primaries had 
been replaced during this molt in any individual. 

In American Kestrel, 92 individuals collected be- 
tween their first January and the initiation of the 
first complete molt had replaced a mean of 90.8% 
of body feathers. The amount of feathers replaced 
ranged from 40% of the body and no wing coverts 
(CAS 73504 collected 26 April) to 100% of the 
body feathers, all lesser coverts, and up to three 
proximal median coverts, but no flight feathers 
(several specimens). 

First-year Molts in other North American Falco- 
niformes. Evidence of body-feather replacement 
prior to initiation of primary molt was recorded in 
16 North American Falconiformes (all but North- 
ern Goshawk and Broad-winged Hawk; Table 1). 
Replaced feathers were observed primarily on the 


back and breast (Fig. 2). Specimens collected in 
fall and early winter had replaced some feathers 
on the crown, throat, and upper back; whereas 
many spring specimens had replaced larger scap- 
ulars and some breast feathers, but had retained 
most to all feathers of the crown, upper back, and 
throat. Newly-replaced feathers on the underparts 
tended to show patterns resembling Juvenal feath- 
ers when molted in fall and definitive feathers 
when molted in spring, with a clinal rate of pattern 
change with time (Fig. 1). The seasonal timing at 
which definitive characteristics in these feathers 
were acquired appeared to vary, occurring during 
the late fall and winter in some species (e.g., Sharp- 
shinned Hawk, Red-shouldered Hawk, and Pere- 
grine Falcon) and during the spring or later in oth- 
ers (e.g., Swainson’s and Rough-legged hawks). 

Extent of this molt showed substantial intraspe- 
cific and interspecific variation (Table 1). Among 
the 16 species showing these molts, the mean max- 
imum recorded extent for all species combined 
was 25.6%, varying from 5% in Golden Eagle to 
50% in Northern Harrier and Merlin (Table 1). 
Examples of specimens showing extensive body- 
feather replacement included Northern Harrier 
MVZ144731 (collected in March with 35% replace- 
ment), Sharp-shinned Hawk MVZ99723 (May, 
45%), Swainson’s Hawk CAS13889 (April, 40%), 
Red-tailed Hawk CAS27181 (April, 20%), Rough- 
legged Hawk MVZ173433 (December, 20%), and 
Peregrine Falcon CAS73587 (December, 25%; Fig. 
2). On the other hand, at least some individuals 
(11-80%) of 12 species (or 11—100% of 14 species 
when Northern Goshawk and Broad-winged Hawk 
are included) collected in March-May showed no 
feather replacement (Table 1), and six specimens 
were recorded that had begun shedding primaries 
during the first complete molt, but remained in 
complete Juvenal body plumage. No first-year birds 
of these 22 species showed symmetrical replace- 
ment of any flight feathers prior to initiation of the 
first complete molt. 

Two replacement patterns according to season 
were observed among these 16 species (Table 1, 
Fig. 3). In Northern Harrier, Sharp-shinned, Coo- 
per’s, Gray, and Ferruginous hawks. Merlin, and 
Peregrine and Prairie falcons, molt had com- 
menced in few birds in fall, some birds in winter, 
and most birds in spring. In Osprey and Red-shoul- 
dered, Red-tailed, and Rough-legged hawks, molt 
had occurred in some birds in fall, appeared to be 
suspended in many birds over winter, and was re- 


382 


Pyle 


VoL. 39, No. 4 



Figure 2. Dorsal and ventral aspects of a first-year Peregrine Falcon {F. p. anatum) collected 15 December in Cali- 
fornia (CAS73587) showing evidence of 25% body-feather replacement during a preformative molt (see Discussion). 
Formative feathers on the dorsum are bluish and barred, and formadve feathers on the underparts are unstreaked 
(breast) and barred (belly and flanks) , resembling definitive feathers in these regions. 


sumed or initiated in some birds in spring. The 
patterns for Bald and Golden eagles, Swainson’s 
Hawk, and Gyrfalcon appeared to be intermediate, 
with molt occurring throughout the first year (Ta- 
ble 1). 

Feather-replacement Patterns in Second-year 
Birds. Totals of 27 birds collected while undergo- 
ing the first complete molt and 146 second-year 
(12-24 mo-old) birds following completion of this 
molt were examined, including at least six individ- 
uals of all species in Table 1 , except for Gray Hawk, 
Red-shouldered Hawk, and Merlin, for which sec- 
ond-year and older individuals could not be distin- 
guished. For the smaller species (including all fal- 


cons) , there was no evidence that feathers replaced 
during the first fall, winter, or spring had been re- 
tained during the first complete molt. All 27 birds 
undergoing this complete molt appeared to be re- 
placing all body feathers (or most feathers in the 
case of the larger species; see below). On second- 
year birds that had completed body molt, all scap- 
ulars as well as crown, back, and underpart feathers 
were uniform in wear, reflecting a complete molt 
within a relatively short time period. For the three 
species mentioned above, examination of 75 birds 
in definitive plumage (likely including second-year 
birds) also showed no variation in feather wear. For 
five of the larger species, Osprey, Bald and Golden 


December 2005 


First-cycle Molts in Falconiformes 


383 



Month 

Figure 3. Proportion of individuals evincing body molt, 
by month, in two species of North American Falconifor- 
mes showing differing seasonal-replacement strategies. 

eagles, and Red-tailed and Ferruginous hawks, var- 
iation in wear among the body feathers precluded 
confirmation that feathers replaced during the first 
year were replaced again during the first complete 
molt. In the two species of eagle, up to many Ju- 
venal body feathers could be retained during the 
first “complete” molt, so it is possible that feathers 
replaced during the first year may also have been 
retained during this molt. 

Discussion 

Traditionally, both the partial first-fall molts of 
North American Kestrels and kites and the com- 
plete molts of other 1-yr-old Falconiformes have 
been considered the “first prebasic molts,” accord- 
ing to the molt terminology of Humphrey and 
Parkes (1959). However, Howell and Corben 
(2000) suggested that the complete first prebasic 
molts of most Falconiformes may be homologous 
with the second prebasic molts of kestrels, kites, 
and most other birds. Accordingly, Howell et al. 
(2003) proposed a revised molt terminology for 
the first cycle, redefining the prejuvenal molt as syn- 
onymous with the first prebasic molt, the complete 
molt at the end of the first molt cycle as the second 
prebasic molt, and any inserted molts during the first 
cycle as preformative molts. 

Evidence of body-feather replacement during 
the first year was found in 18 of 20 species exam- 
ined and, for at least 15 species, the “first complete 
molt” appeared to include those feathers replaced 
during the first cycle. By definition (Humphrey 
and Parkes 1959, Howell et al. 2003), the addition- 
al body-feather replacement during the first year 
should be considered separate, inserted molts; us- 


ing the terminology of Howell et al. (2003), they 
should be considered preformative molts followed 
by the complete or near-complete second prebasic 
molt at a year of age. Under the terminology of 
Humphrey and Parkes (1959), the “first prebasic 
molt” referred to the limited body-feather molt 
during the first fall, winter, or spring of some in- 
dividuals or species and to the complete molt at a 
year of age in other individuals, even within species 
(cf. Elanus caeruleus in Marchant and Higgins 
1993), and presumed homology between individ- 
uals and species was lost. 

In the remaining two species. Northern Gos- 
hawk and Broad-winged Hawk, evidence of prefor- 
mative molts may have been missed in this study 
due to low sample sizes, especially in spring. In- 
deed, body-feather replacement during the first 
year has been reported for Northern Goshawk 
(Bent 1937, Dement’ ev and Gladkov 1951, Fors- 
man 1999). A first-year Broad-winged Hawk (Slater 
Museum of Natural History No. 2367) collected 1 
June reportedly had replaced some breast feathers 
(Clark and Anderson 1984), although this individ- 
ual had initiated the second prebasic molt (D. 
Paulson pers. comm.). It is possible that the con- 
straints of migration may preclude the occurrence 
of a preformative molt in Broad-winged Hawk; 
however, substantial evidence of this molt was 
found in the migratory Swainson’s Hawk. In gen- 
eral, birds that migrate to the tropics or Southern 
Hemisphere often display more extensive first-win- 
ter molts, possibly due to more abundant and sta- 
ble food resources and greater day-lengths with 
which to forage (Myers et al. 1985, Pyle 1998). 
Thus, preformative molts should be expected in 
some Broad-winged Hawks. The evidence, there- 
fore, suggests that preformative molts likely occur 
in at least some individuals of all North American 
Falconiformes. 

Including the pattern of White-tailed Kite and 
American Kestrel, three seasonal strategies of pre- 
formative molt were identified. Consideration of 
the ecology and life history of the species com- 
prising each group revealed no evident explana- 
tions for conditions or constraints leading to each 
molt pattern. There appeared to be a slight phy- 
logenetic component, with a majority of species 
among Accipiter and Falco delaying preformative 
molts until winter or spring (see Herremans and 
Louette 2000) , whereas more species among Buteo 
initiated preformative molts in fall. Northerly 
breeding and wintering species also tended to 


384 


Pyle 


VoL. 39, No. 4 


show a greater amount of preformative molt in fall; 
perhaps first-year birds of these species can take 
advantage of abundant food resources in the fall, 
but suspend molting during winter when food be- 
comes scarce. Such variation might also be expect- 
ed within species that show a wide latitudinal 
breeding range. Within genera, smaller species (in- 
cluding Mdiite-tailed Kite and American Kestrel) 
generally showed higher proportions of birds molt- 
ing a greater amount of feathers. This correlation 
is expected based on the added energy required 
to replace larger feathers (Lindstrom et al. 1993) , 
and may also be part of a signaling mechanism for 
species more likely to undergo breeding in their 
first year (Kemp 1999, Herremans and Louette 
2000). Finally, species inhabiting open areas ex- 
posed to higher amounts of UV radiation appeared 
to undergo more extensive preformative molts. 
However, there were exceptions to all of these pat- 
terns, and it is likely that the extent and timing of 
preformative molts in Falconiformes reflect various 
constraints according to a complex combination of 
species-specific, breeding, migrating, and foraging 
strategies. 

Replacement patterns by season indicate that 
breast feathers and scapulars can be molted later 
in winter or spring during preformative molts, 
while juvenal crown, throat, and upper back feath- 
ers had been retained, an unusual sequence of 
feather replacement for Falconiformes (Palmer 
1988, Wheeler 2003). This suggests a triggering 
mechanism for the preformative molt, by which 
body molt of some feathers is bypassed, resulting 
in spring birds initiating molt at a later point in 
the sequence. Thus, sequence as well as extent may 
vary individually, depending on initiation date. 
Similarly, hormonal processes controlling feather 
pattern appear to develop clinally throughout the 
first year (and sometimes beyond), resulting in de- 
layed acquisition of definitive plumage that varies 
m timing by species. In Swainson’s Hawk and 
Crested Caracara ( Caracara cheriway ) , acquisition of 
definitive plumage is delayed until the following 
summer or later, resulting in identifiable second- 
basic plumages, intermediate in pattern between 
juvenal and definitive plumages (Wheeler 2003). 
For the two eagle species, acquisition of definitive 
plumage requires up to 4 or 5 yr to complete. 

Within many species, the preformative molt ap- 
pears to occur in only a proportion of individuals, 
with some birds retaining full Juvenal plumage un- 
til the second prebasic molt. Similar variation in 


the preformative molt was found among Eurasian 
Accipiters (Herremans and Louette 2000). Thus, 
some individuals exhibit a “Simple Basic Strategy” 
(lacking a preformative molt) whereas others ex- 
hibit a “Complex Basic Strategy ” (including a pre- 
formative molt, but lacking prealternate molts) ac- 
cording to Howell et al. (2003). However, at the 
species level, Falconiformes are best described as 
exhibiting the Complex Basic Strategy, with the 
preformative molt ranging in extent from absent 
to partial. 

Evolutionarily, the existence of these variable 
and at times ephemeral preformative molts may 
suggest that ancestral Falconiformes exhibited 
more extensive preformative molts that have be- 
come vestigial in most (larger) species that do not 
breed at age one and can commence the second 
prebasic molt at an earlier age (Wheeler 2003). 
Phylogenetic evidence suggests that ancestral Fal- 
coniformes branched from a common ancestor 
that also included Podicipediformes, Pelecanifor- 
mes, and Ciconiformes (Sibley and Ahlquist 1990), 
orders which currently display extensive prefor- 
mative molts (Palmer 1962, Howell et al. 2003). 
Alternatively, it is possible that these preformative 
molts in Falconiformes have become inserted over 
time, from an ancestral species that lacked such 
molts; however, the apparent lack of functionality 
for these molts may argue against this alternative 
hypothesis. A better understanding of molt and 
plumage homologies in Falconiformes awaits fur- 
ther study of both molts and phylogenetic relation- 
ships. 

Acknowledgments 

I thank Jack Dumbacher and Douglas J. Long (CAS), 
Carla Cicero (MVZ), and James Dean (USNM) for access 
to collections under their care and Dennis Paulson of the 
Slater Museum of Natural History, Puget Sound, WA, for 
examining a specimen of Broad-winged Hawk for me. 
The manuscript benefited from comments by Steve N.G. 
Howell, Christopher W. Thompson, William S. Clark, Jo- 
sef K. Schmutz, and Marc Herremans. Funding for the 
study of molts and plumages was provided by the Neo- 
tropical Migratory Bird Conservation Act (grant No 
2601) administered by the U.S. Fish and Wildlife Service. 
This is Contribution No. 232 of The Institute for Bird 
Populations. 

Literature Cited 

Bent, A.C. 1937. Life histories of North American rap- 
tors. Part 1. U.S. Nat. Mus. Bull. 167:1-409. 

. 1938. Life histories of North American raptors. 

Part 2. U.S. Natl. Mus. Bull. 170:1-482. 

Clark, W.S. and C.M. Anderson. 1984. First specimen 


December 2005 


First-cycle Molts in Falconiformes 


385 


record of the Broad-winged Hawk for Washington. 
Murrelet 65:93-94. 

Cramp, S. and K.E.L. Simmons. 1980. The birds of the 
western Palearctic. Vol. II. Oxford University Press, 
Oxford, England. 

Dement’ev, G.P. and N.A. Gladkov. 1951. Birds of the 
Soviet Union. Vol. I. Gosudarstvennoe Izdatel’stvo 
“Sovetskaya Nauka,” Moscow, Russia. 

Dickey, D.R. and A.J. van Rossem. 1938. The birds of El 
Salvador. Field Mus. Nat. Hist. Publ. Zool. Sen 23:1-609. 

Eorsman, D. 1999. The raptors of Europe and the Middle 
East. T. and A.D. Poyser, Ltd., London, England. 

Hammerstrom, E. 1967. On the use of fault bars in ageing 
birds of prey. Inland Bird-Banding News 39:35-41. 

Herremans, M. and M. Louette. 2000. A partial post- 
juvenile molt and transitional plumage in the Shikra 
{Acdpiter badius) and Grey Erog Hawk {Accipiter soloen- 
sis).J. Raptor Res. 34:249-261. 

Howell, S.N.G. and G. Corben. 2000. A commentary on 
molt and plumage terminology: implications from the 
Western Gull. West. Birds 31:50-56. 

, C. Corben, P. Pvle, and D.I. Rogers. 2003. The 

first basic problem: A review of molt and plumage 
homologies. Cowdor 105:635-653. 

Humphrey, P.S. and K.C. Parkes. 1959. An approach to 
the study of molts and plumages. Auk 76:1-31. 

Kemp, A.C. 1999. Plumage development and visual com- 
munication in the Greater Kestrel (Falco rupicoloides) 
near Pretoria, South Africa. Ostrich 70:220—224. 

Lindstrom, a, G.H. Visser, and S. Daan. 1993. The en- 
ergetic costs of feather synthesis is proportional to 
basal metabolic rate. Phys. Zool. 66:490-510. 


Marchant, S. and P.J. Higgins. 1993. Handbook of Aus- 
tralian, New Zealand, and Antarctic Birds. Vol. 2. Ox- 
ford University Press, Oxford, England. 

Miller, A.H. 1941. The significance of molt centers 
among secondary remiges in the Ealconiformes, Con- 
dor 43:113-115. 

Miller, K.E, and J.A. Smaixwood. 1997. Juvenal plum- 
age characteristics of male southeastern American 
Kestrels {Falco sparverius paulus) . J. Raptor Res. 31:273- 
274. 

Myers, J.P.,J.L. Maron, and M. Sallaberry. 1985. Going 
to extremes: why do Sanderlings migrate to the trop- 
ics? Ornithol. Monogr. 36:520—535. 

Palmer, R.S. 1962. Handbook of North American birds. 
Vol. 1. Loons through flamingos. Yale University 
Press, New Haven, CT U.S.A. 

. 1988. Handbook of North American birds. Vols. 

4 and 5. Diurnal raptors (Parts 1 and 2). Yale Univer- 
sity Press, New Haven, CT U.S.A. 

Parkes, K.C. 1955. Notes on the molts and plumages of 
the Sparrow Hawk. Wilson Bull. 67:194—199. 

Pyle, P. 1998. Eccentric first-year molt patterns in certain 
tyrannid flycatchers. West. Birds 29:29-35. 

Sibley, C.G. and J.E. Ahlquist. 1990. Phylogeny and clas- 
sification of birds. Yale University Press, New Haven, 
CT U.S.A. 

Wheeler, B.K. 2003. Raptors of western North America. 
Princeton University Press, Princeton, NJ U.S.A. 

Received 12 Eebruary 2004; accepted 6 June 2005 

Associate Editor: Ian G. Warkentin 


J Raptor Res. 39(4) :386— 393 
© 2005 The Raptor Research Foundation, Inc. 


MORPHOMETRIC ANALYSIS OF LARGE FALCO SPECIES AND 
THEIR HYBRIDS WITH IMPLICATIONS FOR CONSERVATION 

Chris P. Eastham^ 

National Avian Research Centre, The Falcon Facility, Penllynin Farm, College Road, 

Carmarthen, Wales SA33 5EH United Kingdom 

Mike K. Nicholes^ 

Ecology Research Group, Canterbury Christ Church University College, Canterbury, Kent, CTl IQU United Kingdom 

Abstract. — Morphometric examination of several large falcon species and their hybrids was conducted to 
ascertain whether phenotype was an accurate indicator of hybrid parentage. Six external body measure- 
ments were recorded from 167 Gyrfalcons {Falco rusticolus), Saker {F. cherrug). Peregrine (F. peregrinus), 
and New Zealand Falcons {E novaezeelandiae) and from 100 FI, F2, and backcross hybrids of these species. 
Principal Component Analysis separated pure species and also indicated clusters for FI peregrine X saker, 
gyr X peregrine, and gyr X saker hybrids. Gyr X peregrine hybrids were distinguishable from their parent 
species, but it was impossible to discriminate accurately between a complex (FI, F2, and backcross) of gyr 
X saker hybrids and between these and the parent species. Escaped or released falconry hybrids are 
perceived as a significant threat to the conservation of wild falcon populations. Under current legislation, 
gyrs and their hybrids are CITES Appendix I species, and sakers are Appendix II species. We suggest that 
phenotypic characteristics are not reliable for identification of such hybrids for legal purposes. Further- 
more, analysis of measurements also identified a “paternal effect,” whereby Fj hybrids, irrespective of 
gender, were phenotypically more similar to their paternal than their maternal progenitors. 

Keywords: Peregrine Falcon', Falco peregrinus; Gyrfalcon-, Falco rusticolus; Saker Falcon-, Falco cherrug; New 

Zealand Falcon', Falco novaezeelandiae; Falcon hybrids, morphometric, principal component analysis:, PCA; CITES. 


anAlisis morfometrico de las especies de ealco de tamano grande y de sus 

HIBRIDOS, E IMPLICACIONES PARA LA CONSERVACION 

Resumen. — Se analizo la morfologia de varias especies de halcones y de sus hfbridos para averiguar si 
el fenotipo es un indicador precise de la paternidad de los hfbridos. Se registraron seis medidas cor- 
porales para un total de 167 individuos pertenecientes a las especies Ealco rusticolus, F. cherrug, E peregrinus 
y E novaezeelandiae, y para un total de 100 hfbridos FI, F2 y retrocruces de estas especies. Un analisis 
de componentes principales separo a las especies puras e identified grupos formados por hfbridos FI 
E peregrinus X E. cherrug, E rusticolus X E peregrinus y E. rusticolus X E cherrug. Los hfbridos F. rusticolus X 
F. peregrinus se diferenciaron de las especies parentales, pero fue imposible distinguir claramente entre 
un complejo (FI, F2, retrocruces) de hfbridos F. rusticolus X F. cherrug, y entre este complejo y las 
especies parentales. Los halcones hfbridos de cetrerfa que escapan o son liberados se consideran una 
amenaza para la conservacidn de las poblaciones silvestres. Bajo la actual legislacidn, F. rusticolus y sus 
hfbridos estan registradas en el Apendice I de CITES y F. cherrug en el Apendice II. Consideramos que 
las caracterfsticas fenotfpicas no son confiables para la identificacidn de estos hfbridos con propdsitos 
legales. Ademas, el analisis morfometrico identified “efectos paternos,” en donde los hfbridos FI, in- 
dependientemente de su sexo, fueron fenotfpicamente mas similares a sus progenitores paternos que 
a los maternos. 

[Traduccidn del equipo editorial] 


^ Present address: Klumpstugevagen 1, 61892 Kolmarden, 
Sweden. 

^ Corresponding author’s address: LEAP, University of 
Greenwich at Medway, Chatham Maritime, Kent ME4 
4TB, United Kingdom. Email: M.Nicholls@gre.ac.uk 


One of the first domestic hybrid falcons was pro- 
duced in 1971 from a female Saker Falcon {Falco 
cherrug and male Peregrine Falcon {F. peregrinus, 
Morris and Stevens 1971, Morris 1972). Since then, 
falconers and raptor breeders have produced many 
different hybrids from members of the Falconifor- 


386 


December 2005 


Morphometrics oe Falcon Hybrids 


387 


Table 1. Identity of hybrid falcons used in the analysis. 


Hybrid Identity^ 

Samples Sizes 

Male Parent 

Fkmat.f. Parent 

FIs 




Gyr X Peregrine 

6 d, 7 9 

Gyr 

Peregrine 

Gyr X Saker 

7 c?, 5 9 

Gyr 

Saker 

Peregrine X Saker 

3 c?, 13 9 

Peregrine 

Saker 

Peregrine X Gyr 

1 c? 

Peregrine 

Gyr 

Peregrine X New Zealand 

1 d 

Peregrine 

New Zealand 

Gyr/Saker X Peregrine 

1 d, 1 9 

Gyr X Saker F2 hybrid 

Peregrine 

F2s 




Gyr X Saker 

4 d, 5 9 

Gyr X Saker FI hybrid 

Gyr X Saker FI hybrid 

Backcrosses — 1st generation 




Gyr X Gyr/Saker 

3 d, 1 9 

Gyr 

Gyr X Saker FI hybrid 

Saker X Gyr/Saker 

3 d, 3 9 

Saker 

Gyr X Saker FI hybrid 

Gyr/Saker X Saker 

14 d, 17 9 

Gyr X Saker FI hybrid 

Saker 

Gyr/Peregrine X Peregrine 

2 d 

Gyr X Peregrine FI hybrid 

Peregrine 

Backcrosses — 2nd generation 




Gyr (3/8)VSaker 

1 d 

Gyr X Saker FI hybrid 

Gyr/Saker X Saker 




(backcross hybrid) 

Gyr (5/8)VSaker 

1 d, 1 9 

Gyr 

Gyr/Saker X Saker 




(backcross hybrid) 


® When naming hybrids, the paternal species is cited first. For example, a cross between a male gyr and female saker is a gyr X saker 
(or gyr/ saker) hybrid, whereas a male saker crossed with a female gyr is a saker X gyr (saker/ gyr) hybrid. 

These numbers represent a simple way to show the proportion of genes from the parent species, assuming that a FI hybrid inherits 
Vz of the genes from both the male and female parent species. For example, a gyr (%) /saker, produced by backcrossing a gyr X saker 
FI hybrid with a gyr/ saker backcross hybrid, has \ gyrfalcon and \ saker genes. 


mes (Boyd and Boyd 1975, Cade and Weaver 1976, 
Bunnell 1986, Weaver and Cade 1991), including 
intergeneric hybrids (e.g., Harris’s Hawks {Para- 
buteo unicinctus] X Cooper’s Hawk [Aedpiter coope- 
ni\ and Harris’s Hawk X Ferruginous Hawk {Buteo 
regalis] ; Fox and Sherrod 1999a) for falconry pur- 
poses. 

Many Fj hybrids are fully viable (Heidenreich 
1997), in their turn producing Fg hybrids or back- 
crosses (Bj and B^ representing 1st and 2nd gen- 
eration backcrosses) to one or other parent spe- 
cies. Indeed, hybrids from within the subgenus 
Hierofalco, the “desert falcon” group (Heidenreich 
1997), exhibit full fertility, presumably over indef- 
inite generations. Less closely-related pairs of spe- 
cies, such as gyr (F. rusticolus) and peregrine, pro- 
duce hybrids with reduced fertility, manifest as 
deformed spermatozoa, completely sterile females 
(Heidenreich and Kuspert 1992), or unviable em- 
bryos (Rosenkranz 1995). 

This extended viability of some falcon hybrids 
coupled with increasing demand over the last 10 
yr for domestic falcons from North American, Eu- 


ropean, and Arabian falconry markets (Fox and 
Sherrod 1999b) has prompted conservation con- 
cerns. Escaped domestic hybrids may be merely a 
curiosity (Forseman 1999), a nuisance for bird 
watchers (Gantlett and Millington 1992), or a 
threat to the integrity of wild populations. Indeed, 
falcon pairs made up of an escaped hybrid and a 
wild, pure individual have been documented sev- 
eral times (e.g., Kleinstauber and Seeber 2000, 
Lindberg 2000) . A further conservation issue pre- 
sumably concerns illegal-trade in falcons, whereby 
protected falcon species may be “laundered” as 
domestic hybrids. In this study, we examine the re- 
lationship between falcon species and their hy- 
brids, particularly the accuracy of using morpho- 
metric characters for identification, and discuss the 
conservation issues concerning falcon hybrids. 

Methods 

We investigated four large falcon species, namely Per- 
egrine, Gyr, Saker, and New Zealand falcons {F. novaezee- 
landiae) and several of their hybrid types used for falcon- 
ry. Hybrid falcons were all bred in captivity and, 
therefore, their parentage was known. Six external body 


388 


Eastham and Nicholls 


VoL. 39, No. 4 


Table 2. Principal Component Analysis (PCA) of six anatomical measurements from juvenile male Gyrfalcon, Saker, 
Peregrine, and New Zealand falcon species and hybrids of those species. Eigenvalues and eigenvectors (based on the 
correlation matrix). 




Principai. Component 


1 

2 

3 

4 

Eigenvalue 

3.0328 

1.3691 

0.7459 

0.6833 

Percent of variability 

0.5055 

0.2282 

0.1243 

0.1139 

Cumulated Percent 

0.5055 

0.7336 

0.8580 

0.9718 

Characters 



Eigenvectors 


Wing chord 

0.4990 

-0.2324 

0.1490 

-0.3852 

Wing width 

0.5552 

-0.0321 

-0.0688 

-0.1697 

Tail length 

0.5195 

0.2308 

-0.2347 

-0.1471 

Tail step 

0.0247 

0.6402 

0.7536 

-0.1395 

Tarsus length 

0.3536 

0.3591 

-0.1598 

0.7769 

Digit three length 

0.2170 

-0.5940 

0.5697 

0.4221 


measurements were collected from live juvenile Gyrfal- 
cons {N = 7 males and 6 females) , Saker {N = 34 males 
and 40 females). Peregrine {N = 17 males and 24 fe- 
males), and New Zealand falcons {N = 25 males and 14 
females), and their various hybrids (Table 1). Apart from 
European Peregrine Falcon subspecies, the majority of 
which were F. peregrinus peregrinus, no other differentia- 
tion was made between subspecies or geographic 
morphs. All birds were kept at the National Avian Re- 
search Center’s Falcon Facility in Carmarthen, Wales, 
U.K. The majority of Fg and backcross 1 and 2 hybrids 
were between gyrs and sakers. This is because hybrids 
between members of the subgenus Hierofalco remain fer- 
tile for an indefinite number of generations, whereas hy- 
brids between more out-crossed falcon species, such as 
peregrines and sakers, have a reduced fertility. Some of 
the hybrids included are produced in very low numbers 
(e.g., Peregrine Falcon X New Zealand Falcon), and pub- 
lished data on these are rare. Therefore, we included 
them in the analysis. 

One of us (C. Eastham) took six measurements, name- 
ly wing chord length and width, tail length, tail step (the 
difference between the outermost tail feather [rectrix 6] 
and the tip of the center tail feather [rectrix 1] on the 
same side), tarsus length, and third digit length from 
each bird. Measurement protocols followed standard 
methods described by Baldwin et al. (1931), Fox (1977), 
Biggs et al. (1978), Kemp (1987), and Fox et al. (1997). 
Feather characters were measured to the nearest 1 mm 
and non-feather characters to the nearest 0.1 mm using 
a pair of digital calipers, a steel ruler, and tape measure. 
Inclusion of single individuals, for example, a male Per- 
egrine Falcon X New Zealand Falcon, allowed us to em- 
ploy Principal Component Analysis (PCA) on XLSTAT- 
Pro (Fahmy 1998) statistical software as a suitable method 
for data analysis. Male and female data were analyzed 
separately to eliminate background variation due to re- 
versed sexual size dimorphism (Brown and Amadon 
1968, Cade 1982), or other sex-linked or sex-limited char- 
acters. We used wing chord length to distinguish males 


and females, as this measurement was reported by Wyllie 
and Newton (1994) and Eastham (2000) to be the most 
reliable indicator of overall body size. 

Results 

Males. Principal Component (PC) 1 (Table 2, 
Fig. 1) accounted for the majority (50.5%) of var- 
iation. Because all eigenvectors for PCI showed 
positive and nearly equal values, we concluded this 
component represents overall body size (Wiley 
1981). Male gyrs and FI and F2 gyr/saker hybrids 
had the largest body size, whilst male peregrines 
and New Zealand Falcons were the smallest. 

PC 2 (Table 2, Fig. 1) accounted for 22.8% of 
the variation, as indicated by a contrast in eigen- 
vectors between the positively weighted tail step, 
tail and tarsus length, the negatively weighted digit 
three length, and wing chord length and width 
(Table 2). This component summarizes variation 
related to body shape. Tail step and digit three 
length showed the strongest positive and negative 
weightings, respectively. With a low negative 
weighting, wing width was of limited use in further 
analysis of PC 2. New Zealand Falcons had the rel- 
atively longest tail step (indicating a more rounded 
tail) , tail and tarsus length, and shortest digit three 
and wing chord, whilst peregrines, the single per- 
egrine X New Zealand hybrid and gyr/ saker X per- 
egrine hybrid had the relatively longest digit three 
and wing chord length and shortest tail step and 
tail and tarsus length. FI gyr X peregrines and sa- 
kers showed a wide variation in PC 2 values, with 
an individual saker having the highest PC 2 value. 


December 2005 


Morphometrics of Falcon Hybrids 


389 



Principal Component 1 

• Gyrfalcon BSaker JKPeregrine -New Zealand falcon AGyr/ Peregrine +Gyr/Saker- FI • Gyr / Saker - F2 

X Peregrine / Saker A Peregrine / New Zealand ^ Gyr x Gyr/ Saker □ Gyr/ Saker x Saker OSakerx Gyr/ Saker :: Gyr (5/8) / Saker SSGyr/ Saker x Peregrine 


Figure 1. Principal component scores from morphometric comparison of various male falcon species and their 
hybrids. 


PC 3 and 4 (Table 2) accounted for only 12.4% 
and 11.3% of the residual variation, respectively. 
Tail step and digit three length had a high positive 
weighting in PC 3, and in PC 4, there was a con- 
trast between positively weighted tarsus and digit 
three length and negatively weighted wing chord 
length. As PC 1 and 2 together accounted for the 
majority (73%) of variation, we did not consider 
PC 3 and 4 further. 

Females. The PCA for juvenile female falcons 
showed a similar pattern of variation as that seen 
in juvenile males. Again PC 1 (Table 3, Fig. 2) ac- 
counted for the largest proportion (44.7%) of var- 
iation and as indicated by mostly positive and near- 
ly equal values represents, as with males, overall 
size (Wiley 1981). Gyrs and the various gyr/ saker 
FI, F2 and backcrosses had the largest size, whilst 
Peregrine and New Zealand falcons were the small- 
est. 

PC 2 (Table 3, Fig. 2) accounted for a further 
21.8% of the variation, and we concluded that this, 
again like males, was related to shape. This was in- 
dicated by a contrast in eigenvectors between tbe 


positively weighted tail step, tail and tarsus length, 
the negatively weighted digit three length, and 
wing chord length and width (Table 3). Positively 
weighted tail step and negatively weighted digit 
three and wing chord length had the highest ei- 
genvectors for this PC. New Zealand Falcons had 
the longest tail step and the shortest digit three 
and wing chord length, whilst FI gyr X peregrines 
and peregrines had the shortest tail length and the 
longest digit three and wing chord length. PC 3 
and PC 4 (Table 3) accounted for 18.4% and 
11.4% of the variation, respectively. As for males, 
we did not consider these principal components 
further. 

Discussion 

Using PCA we found that the four falcon species, 
irrespective of sex, were clearly separated into 
groups: New Zealand Falcons with a small size, 
long rounded tails and tarsi, and short wings; per- 
egrines, also with a small size, long digit three 
lengths, and long narrow wings; sakers with a large 


390 


Eastham and Nicholls 


VOL. 39, No. 4 


Table 3. Principal Component Analysis (PCA) of six anatomical measurements from juvenile female Gyrfalcon, 
Saker, Peregrine, and New Zealand falcon species and hybrids of those species. Eigenvalues and eigenvectors (based 
on the correlation matrix). 


Principal Component 



1 

2 

3 

4 

Eigenvalue 

2.6826 

1.3071 

1.1023 

0.6864 

Percent of variability 

0.4471 

0.2179 

0.1837 

0.1144 

Cumulated percent 

0.4471 

0.6649 

0.8487 

0.9631 

Characters 


Eigenvectors 



Wing chord 

0.4717 

-0.3853 

■0.1963 

0.4118 

Wing width 

0.5765 

-0.1767 

■0.0356 

-0.0129 

Tail length 

0.5737 

0.1670 

■0.0642 

-0.1107 

Tail step 

0.0866 

0.6865 

0.1871 

0.6869 

Tarsus length 

0.3077 

0.2169 

0.6899 

-0.4484 

Digit three length 

-0.1177 

-0.5236 

0.6672 

0.3809 


size, long rounded tails, and short digits; and gyrs 
with the largest size. 

Using these external body measurements, we 
also found it possible to identify three main hybrid 
groups: a complex of FI, F2 and backcross gyr/ 
sakers; FI gyr X peregrines and FI peregrine X 
sakers. Further, it was possible to separate gyr X 
peregrines from their parent species, but impossi- 
ble to separate completely the FI, F2, and back- 
cross gyr/ sakers hybrid-complex from pure sakers 
or particularly, from pure gyrs. 

Overall, we found that the hybrids were gener- 
ally of intermediate phenotype between their par- 
ents. However, beyond this it appears that the pa- 
ternal progenitor influences the phenotype to a 
greater extent than maternal. For example, for 
both males and females, the clusters representing 
FI gyr X peregrine hybrids, hybrids whose male 
parents were gyrs, were spatially closer to the gyr 
clusters than to the peregrine clusters (Fig. 1, 2). 
Thus, both male and female gyr X peregrine hy- 
brids have a morphology closer to that of gyrs than 
to peregrines. Further, that the single female per- 
egrine X gyr, whose sire was a peregrine, was spa- 
tially closer to the peregrine cluster than the gyr 
cluster, adds further weight to this proposed gen- 
erality. Similarly, the male and female FI gyr X sa- 
kers, both with gyrs as male parent, appear more 
gyr-like than saker-like in morphology. Except for 
male FI peregrine X sakers this “paternal effect” 
seems true for all species combinations. We ex- 
plain this by considering that two thirds of the sex 
linked genes in a population are carried by the 


homogametic sex (male in birds; Mittwoch 1977) 
and only one third by the heterogametic sex (Fal- 
coner 1967). Therefore, falcon sires (the homo- 
gametic sex) will contribute more sex-linked alleles 
to their hybrid offspring than will (heterogametic) 
dams. 

International trade in endangered species, such 
as some falcons, can be a profitable enterprise and, 
if unregulated, can threaten their conservation. 
Regulation of the trade in such endangered spe- 
cies is by international agreements, such as the 
Convention on International Trade in Endangered 
Species of Wild Fauna and Flora (CITES). Accord- 
ing to their degree of endangerment in the wild, 
all species are classed in one of three CITES ap- 
pendices. Special conditions apply to the most en- 
dangered, known as Appendix I species (i.e., those 
threatened with extinction and whose survival 
could be impaired by trade) , which allows restrict- 
ed trade in captive and domestic bred individuals. 
Appendix II includes species considered less 
threatened. The saker is an Appendix II species, 
and although trade is regulated, this is less exact- 
ing than for Appendix I species. Despite relatively 
healthy world populations and for reasons which 
are largely political (White and Kiff 1998), gyrs and 
peregrines and their hybrids are included in Ap- 
pendix I. 

The results we present here show that it can be 
difficult to discriminate falcon species accurately 
from their hybrids, especially hybrids of Appendix 
I gyrs and Appendix II sakers. Similarly, plumage 
variation, especially in juvenile falcons, is difficult 


December 2005 


Morphometrics of Falcon Hybrids 


391 



• Gyrfalcon 

■ Saker 

)K Peregrine 

- New Zealand falcon 

A Gyr /Peregrine 

+ Gyr /Saker - F1 

• Gyr / Saker - F2 

X Peregrine / Saker 

>I< Peregrine / Gyr 

<>Gyr X Gyr / Saker 

A Gyr / Peregrine x Peregrine 

□ Gyr / Saker x Saker 

O Saker x Gyr / Saker 

:: Gyr (5/8) / Saker 

— Gyr (3/8) / Saker 

SSGyr / Saker x Peregrine 




Figure 2. Principal component scores from morphometric comparison of various female falcon species and their 
hybrids. 


to assess objectively and make comparisons be- 
tween species and their hybrids. These observa- 
tions may provide fuel for two separate arguments. 

Ornithologists are increasingly aware of the 
widespread genomic compatibility and potential 
for hybridization amongst what appear to be very 
dissimilar species (Grant and Grant 1992), such 
that hybridization between avian species is consid- 
ered more common than originally thought (Gill 
1998). Mayr and Short (1970) estimated that up to 
10% of North American bird species regularly hy- 
bridize; it’s so common that hybrids are even in- 
cluded in birdwatchers’ field guides (Sibley 2000). 
The presence of natural hybrids is not believed to 
be a threat to the integrity of a species, even 
though they may challenge the biological species 
concept of taxonomists (Brookes 1999). Amongst 
free-living, wild birds of prey such hybridization is 
increasingly documented at the subspecific (Fefe- 
lov 2001), specific (Hamer et al. 1994), and even 
to the intergeneric levels (Corso and Gildi 1998, 
Yosef et al. 2001). 

Introgressive hybridization may therefore be a 
process by which species evolve, rather than some- 


thing that will corrupt them. Thus, if the species 
concept for birds is much looser than conservation 
law dictates, then perhaps it is the legislation and 
not the species concept that must be challenged. 
For example, the so-called “Altai falcon” (Falco al- 
taicus Menzbier) , whose phenotype seems to share 
characters with both gyrs and sakers, is believed by 
some to be the result of introgressive hybridization 
between gyrs and sakers, rendering all these as 
allospecies within a single superspecies (Pfander 
1987, Ellis 1995a, 1995b). This being so, then to 
discriminate between Appendix I gyrs and Appen- 
dix II sakers may be irrelevant, and artificially pro- 
duced crosses between these two may be of no evo- 
lutionary threat should they escape to the wild. If, 
however, the Altai Falcon is merely a large, dark 
race of the saker (Eastham and Nicholls 2002) , and 
any resemblance to gyrs is merely superficial, then 
escaped hybrids between these species could po- 
tentially compromise wild populations, and the in- 
tegrity of gyrs and sakers must be recognized and 
CITES regulations enforced. 

A different view of the role of natural hybridiza- 
tion accepts that avian species are dynamic entities. 


392 


Eastham and Nicholls 


VoL. 39, No. 4 


which in certain circumstances, freely exchange 
genes with other such entities. Despite these inter- 
actions, the integrity of the whole is a fragile one, 
and to short circuit gene flow via artificial hybrids 
is a danger to this integrity. CITES protocols are 
the response to perceived conservation status, and 
therefore, it should be mandatory to discriminate 
between species as we know them. Accurate iden- 
tification to assist in controlling the trade in fal- 
cons is paramount, and we have shown that criteria 
other than phenotypic characteristics (e.g., DNA 
markers) must be employed to identify individuals 
and species. This is imperative if CITES is to re- 
main an effective means of regulating legitimate 
trade and protecting species in the wild. 

Acknowled gments 

Thanks to His Highness Sheikh Khalifa bin Zayed A1 
Nahyan and His Highness Sheikh Hamdan bin Zayed A1 
Nahyan for sponsoring this research. Thanks also to M. 
A1 Bowardi, managing director of the Environmental Re- 
search and Wildlife Development Agency (ERWDA), Drs. 
Nick Fox, Director of the ERWDA Falcon Programme, 
Robert Kenward, and Juan Negro and Violeta Munoz for 
comments on this manuscript. 

Literature Cited 

Batdwin, S.R, H.C. Oberholser, and L.G. Worley. 1931. 
Measurements of birds. Sci, Publ. Clevel. Mus. Nat. Hist. 
2:1-65. 

Biggs, H.C., R. Biggs, and A.C. Kemp. 1978. Measure- 
ments of raptors. Pages 77-82 in A.C. Kemp [Ed.], 
Proceedings of a Symposium on African Predatory 
Birds. Northern Transvaal Ornithological Society, Pre- 
toria, South Africa. 

Boyd, L.L. and N. Boyd. 1975. Hybrid falcons. Hawk 
Chalk 14:53-54. 

Brookes, M. 1999. Live and let live. New Scientist 2193: 
32-36. 

Brown, L. and D. Amadon. 1968. Eagles, hawks and fal- 
cons of the world. Vol. 2. The Wellfleet Press, Secau- 
cus, NJ U.S.A. 

Bunnell, S. 1986. Hybrid falcon overview — 1985. Hawk 
Chalk 25:43-47. 

Burton, J.F. 1995. Birds and climate change. Christopher 
Helm Publishers, London, England. 

Cade, T.J. 1982. Falcons of the world. William Collins 
Sons and Co. Ltd., London, England. 

andJ.D. Weaver. 1976. Gyrfalcon X peregrine hy- 
brids produced by artificial insemination./. North Am. 
Falconers Assoc. 15:42-47. 

CORSO, A. AND R. Gildi. 1998. Hybrids of Black Kite and 
Common Buzzard in Italy in 1996. Dutch Birding 20: 
226-233. 

Eastham, C.P. 2000. Morphological studies of taxonomy 
of the Saker (Falco cherrug; Gray 1833) and closely 


allied species. Ph.D. thesis. University of Kent at Can- 
terbury, Canterbury, Kent, England. 

and M.K. Nicholls. 2002. Morphological classi- 
fication of the so-called “Altai Falcon.” Pages 211-219 
in R. Yosef, M.L. Miller, D. Pepler [Eds.], Raptors in 
the new millennium. Int. Birding and Res. Centre, 
Eilat, Israel. 

Ellis, D.H. 1995a. The Altai falcon: origin, morphology, 
and distribution. Pages 143-168 m Middle East Falcon 
Research Group [Eds.], Proceedings of the Specialist 
Workshop. Abu Dhabi, United Arab Emirates. 

. 1995b. What is Falco altaicus Menzbier? /. Raptor 

Res. 29.45-25. 

Fahmy, T. 1998. XLSTAT-PRO statistical software. Paris, 
France. 

Falconer, D.S. 1967. Introduction to quantitative genet- 
ics. Oliver and Boyd, London, England. 

Fefelov, I.V. 2001. Comparative breeding ecology and 
hybridization of eastern and western Marsh Harriers 
Circus spilonotus and C. aeruginosus in the Baikal region 
of eastern Siberia. Ibis 143:587-592. 

Forseman, R. 1999. The raptors of Europe and the Mid- 
dle East: a handbook for field identification. T. and 
A.D. Poyser, Ltd., London, England. 

Fox, N.C. 1977. The biology of the New Zealand Falcon 
(Falco novaezeelandiae; Gmelin 1788). Ph.D. disserta- 
tion, University of Canterbury, New Zealand. 

, C.P. Eastham, and HJ. MacDonald. 1997. The 

ERWDA handbook of falcon protocols. Environmen- 
tal Research and Wildlife Agency, Abu Dhabi, United 
Arab Emirates. 

and S. Sherrod. 1999a. The use of exotic and 

hybrid raptors in falconry. Information arising from 
the International Committee on Hybrids. Internation- 
al Wildlife Consultants Ltd., Carmarthen, Wales, Eng- 
land. 

AND S. Sherrod. 1999b. Cited in “Scale produc- 
tion and use of hybrids in falconry.” Falco 13, Middle 
East Falcon Research Group, International Wildlife 
Consultants Ltd., Carmarthen, Wales, England. 

Gantlett, S. and R. Millington. 1992. Identification of 
large falcons. Birding World 5:101-106. 

Gill, F.B. 1998. Hybridization in birds. Auk 115:281—283. 

Grant, P.R. and B.R. Grant. 1992. Hybridization of bird 
species. SAmce 256:193-197. 

Hamer, T.E., E.D. Forsman, A.D. Fuchs, and M.L. Wal- 
ters. 1994. Hybridization between Barred and Spot- 
ted owls. Auk 111:487-492. 

Heidenreich, M. 1997. Birds of prey: medicine and man- 
agement. Blackwell Scientific Ltd., Oxford, England. 

AND H. Kuspert. 1992. Metodos de la reproduc- 

cion en cautividad de las diferentes especies de hal- 
cones. El problema de los hibridos y la cooperacion 
entre criadores y autoridades. Proc. Congr. Nac. Cria. 
Aves Cetreria 1:15—17. 

Hoeller, T. AND P. Wegner. 2001. Der Wanderfalke ist 
in Gefahr! Die zweite Hybridfalken-Wildbrut in 


December 2005 


Morphometrics of Falcon Hybrids 


393 


Deutschland zeigt die Brisanz der Hybridfalkenzucht. 
Greifvogel u. Falknerei 2000:64—66, 

Kemp, A.C. 1987. Taxonomy and systematics. Pages 251- 
259 in B.A. Giron Pendleton, B.A. Millsap, K.W. Cline, 
and D.M. Bird [Eds.], Raptor management tech- 
niques manual. National Wildlife Federation Scientif- 
ic and Technical Series No. 10. 

Kleinstauber, G. and H.J. Seeber. 2000. Die erfolgreiche 
Brut eines Gerfalke-X-Wanderfalke-Hybriden {Falco 
rusticolus X Falco peregrinus) in freier Wildbahn-Re- 
port, MaBnahmen, SchluBfolgerungen. Poplationso- 
kol. Greifvogel- u. Eulenarten 4:323-332 

Lindberg, P. 2000. Hybridfalk hackar igen. Var Fagelvdrld 
8:28-29. 

Mayr, E, 1963. Animal species and evolution. Belknap 
Press of Harvard University Press, Cambridge, MA 

U.S.A. 

AND L.L. Short. 1970. Species taxa of North 

American birds. No. 9 Publ. Nuttal Ornithol. Club. 

Mittwoch, U. 1977. Genetics of sex differentiation. Ac- 
ademic Press, Burlington, MA U.S.A. 

Morris, J. 1972. Peregrine/ Saker breeding. Cap. Breed. 
Diurn. Birds of Prey 1:14-15. 

AND R. Stevens. 1971. Successful cross-breeding 

of Peregrine tiercel and a Saker falcon. Cap. Breed. 
Diurn. Birds of Prey, 1:5-7. 

Pennycuick, C.J. 1989. Bird flight performance: a prac- 


tical calculation manual. Oxford University Press, Ox- 
ford, England. 

Pfander, P.V. 1987. Once again concerning the Altay Gyr- 
falcon. Selevinia 2:38-42. 

Rosenkranz, D. 1995. Cited in M. Heidenreich. 1997. 
Birds of prey: medicine and management. Blackwell 
Scientific Ltd., Oxford, England. 

Sibley, C. 2000. The North American bird guide. Chris- 
topher Helm, London, England. 

Weaver, J.D. and T.J. Cade. 1991. Falcon propagation. 
The Peregrine Fund, Inc., Boise, ID U.S.A. 

White, C.M. and L.F. Kiff. 1998. Language use and mis- 
applied, selective “science;” their roles in swaying 
public opinion and policy as shown with two North 
American raptors. Pages 547—560 in R.D. Chancellor, 
B.-U. Meyburg, andJ.J. Ferrero [Eds.], Holarc tic birds 
of prey. ADENEX-WWGBP, Badajoz, Spain. 

Wiley, E.O. 1981. Phylogenetics: the theory and practice 
of phylogenetic systems. Wiley/Interscience, Hobo- 
ken, NJ U.S.A. 

Wyllie, I. AND I. Newton. 1994. Latitudinal variation m 
the body-size of Sparrowhawks {Accipiter nisus) within 
Britain. Ibis 136:434-440. 

Yosef, R., A.J. Helbig, and W.S. Clark. 2001. An intra- 
generic Accipiter hybrid from Eilat, Israel. Sandgrouse 
23:141-144. 

Received 13 April 2004; accepted 24 March 2005 

Former Associate Editor: Juan Jose Negro 


/. Raptor Res. 39(4):394-403 
© 2005 The Raptor Research Foundation, Inc. 

A CHANGE IN FORAGING SUCCESS AND COOPERATIVE 
HUNTING BY A BREEDING PAIR OF PEREGRINE FALCONS 

AND THEIR FLEDGLINGS 

Dick Dekker^ 

3819-112A Street NW, Edmonton, Alberta, T6J 1K4 Canada 

Robert Taylor 

P.O. Box 3105, Spruce Grove, Alberta, T7X 3A1 Canada 

Abstract. — The foraging habits of one pair of Peregrine Falcons (Falco peregrinus) nesting on a power 
plant in central Alberta were studied over seven consecutive breeding seasons (1998-2004) . We observed 
386 attacks that resulted in II7 captures of prey. The success rate increased from 21.9% in the first year 
to 39.1% in the seventh year and averaged 30.3%. The majority of hunts (76.7%) were initiated from 
soaring, and the peregrines commonly used the hot air above the plant’s smoke stacks to gain height. 
The success rates of hunts launched from soaring versus perches were not significantly different (28.7% 
versus 35.6%). Tandem hunts by the pair {N = 100) were more successful than solo hunts (39.0% versus 
27.3%), but the difference was not significant. The main prey species were Franklin’s Gull {Larus pipix- 
can) and small passerines, which made up 53.0% and 27.4% of kills, respectively. There was a significant 
difference in the respective capture rates of these prey types (42.5% versus 24.1%). The success rates 
of the male and the female peregrines were not significantly different, but there was a notable difference 
in the prey taxa taken by each gender. The male captured 84.4% of the passerines, but only 22.9% of 
the gulls. After the first year, there was a significant switch from passerines to gulls, which paralleled a 
significant change in gender participation in foraging. All gulls captured by the male were surrendered 
to the female or to the fledged Juveniles. In 23.3% of hunts, one or both parents were accompanied 
by one or more fledglings. The male participated in 76.7% of all parent-fledgling hunts {N = 90), the 
female in 10.0%, and the remainder by both parents. Aerial prey transfers from adults to flying young 
were feet-to-feet or through aerial drops. The hypothesis that parent peregrines make live-drops of prey 
to their fledged young to teach them how to hunt is confounded by observations that live-drops of just- 
caught prey are also made by the adult male to his mate. However, the hypothesis that peregrines assist 
their young in capturing prey is supported by anecdotal evidence. 

Keywords: Peregrine Falcon-, Falco peregrinus; tandem hunting, fledgling hunting, adult hunting cooperative 

hunting. 


CAMBIOS EN EL EXITO DE FORRAJEO Y CAZA COOPERATIVA EN UNA PAREJA DE FALCO PER- 
EGRINUS Y SUS VOLANTONES 

Resumen. — Se estudiaron los habitos de forrajeo de una pareja de Falco peregrinus que anido en una 
planta de energfa en el centro de Alberta a lo largo de siete epocas reproductivas consecutivas (1998 a 
2004). Observamos 386 ataques, los cuales resultaron en 117 capturas de presas. La tasa de exito se 
incremento del 21.9% en el primer ano al 39.1% en el septimo ano, y en promedio fue del 30.3%. La 
mayoria de los eventos de caza (76.7%) fueron iniciados a partir de vuelos elevados; los halcones 
emplearon frecuentemente el aire caliente que se encontraba encima de las columnas de humo de la 
planta para alcanzar mayores alturas. La tasa de exito de los ataques iniciados en vuelo no fue signifi- 
cativamente diferente de la de los ataques iniciados desde perchas (28.7% versus 35.6%). Los eventos 
de caza en los que ambos miembros de la pareja participaron en tandem {N = 100) fueron mas exitosos 
que aquellos en los que participo un solo individuo (39.0% versus 27.3%), pero la diferencia no fue 
significativa. Las presas predominantes fueron la gaviota Larus pipixcan y varias aves paserinas pequenas, 
representando el 53.0% y el 27.4% de las capturas, respectivamente. Existio una diferencia significativa 
en las tasas de captura de los distintos tipos de presa (42.5% para L. pipixcan y 24.1% para las aves 


^ Email address: tj_dick_dekker@hotmail.com 


394 


December 2005 


Foraging Success of Peregrine Falcons 


395 


paserinas) . Las tasas de exito no fueron diferentes entre el macho y la hembra, pero existio una notable 
diferencia en los taxa capturados por cada individuo. El macho capture el 84.4% de las aves paserinas, 
pero solo el 22.9% de las gaviotas. Despues del primer ano, existio un cambio significativo de paserinas 
a gaviotas de forma paralela con un cambio significativo en la participacion de los miembros de la 
pareja en el forrajeo. Todas las gaviotas capturadas por el macho fueron entregadas a la hembra o a 
los volantones. En el 23.3% de los eventos de caza, uno o los dos padres estuvieron acompahados por 
uno o mas volantones. El macho participo en el 76.7% de todas las cacerias en las que participaron 
volantones {N = 90), la hembra en el 10.0% de estas, y ambos individuos en el porcentaje restante. Los 
adultos entregaron las presas a los juveniles en vuelo directamente de garras a garras, o dejandolas caer 
en el aire. La hipotesis de que los parentales dejan caer presas vivas para ensenarles como cazar a sus 
volantones es dificil de apoyar ya que el macho adulto tambien dejo caer presas recien capturadas en 
vuelo para que su pareja las tomara. Sin embargo, la hipotesis de que los peregrinos le ayudan a sus 
volantones a capturar presas es apoyada por informacion anecdotica. 

[Traduccion del equipo editorial] 


Cooperative or tandem hunting by mated pairs 
of Peregrine Falcons {Falco peregrinus), in which 
both partners attack the same prey simultaneously, 
has been reported from many locations across the 
species’ worldwide range (e.g., Cade 1960, Bird 
and Aubry 1982, Thiollay 1988, Frank 1994, Tre- 
leaven 1998, Jenkins 2000). The success rate of tan- 
dem hunts has been reported to be higher than 
hunts by single peregrines (Thiollay 1988), but 
whether or not tandem hunting by mated pairs ac- 
tually involves a degree of coordination between 
the two individuals is unclear. On their breeding 
range, Peregrine Falcons are also known to hunt 
together with their fledglings, although data are 
limited (Brown and Amadon 1968, Palmer 1988, 
White et al. 2002). The notion that juvenile pere- 
grines need to be taught hunting skills by their par- 
ents is contradicted by the fact that captive-reared 
peregrines, deprived of parental instruction, begin 
pursuing and capturing prey at about the same age 
as young falcons at natural nest sites (Cade and 
Burnham 2003). Apparently, the peregrine does 
not need to be taught how to chase and kill prey 
(Cade 1982, Sherrod 1983). However, Newton 
(1979) and Ratcliffe (1993) state that more critical 
observation is needed during the period in which 
fledglings achieve independence. 

This paper presents empirical data on the for- 
aging behavior of one mated pair of peregrines 
hunting solo or in tandem over seven consecutive 
breeding seasons. Additionally, we present anec- 
dotal observations on parent-fledgling interac- 
tions. 

Study Area and Methods 

The study area is in central Alberta, Canada, at 53°N. 
In this largely agricultural region, peregrines nested com- 
monly on cliffs and cutbanks along rivers and creeks until 


they became extirpated (Dekker 1967, Fyfe 1976). A gov- 
ernment program of releasing captive-reared peregrines 
in central Alberta began in 1975 and led to the establish- 
ment of breeding pairs on city buildings by 1981 (Hol- 
royd and Banasch 1990). By the mid 1990s, peregrines 
began using nest boxes on high industrial structures in 
rural regions. The breeding site selected for this study is 
a power plant on the north shore of Wabamun Lake, 
which is roughly 8 X 20 km in size. The nest box was 
built by the Alberta Falconry Association in 1993 and put 
in place by TransAlta Utilities (Wabamun, Alberta) on a 
catwalk just below the top of a 91 m smoke stack. Lower 
down, the flat roof of the main building functions as a 
landing pad for newly-fledged young. Power line pylons 
in the vicinity provide high perches and plucking posts. 
Settling ponds and bare ground adjacent to the plant are 
used as loafing areas by the fledglings and by the adults 
that bring down large prey such as gulls and ducks. The 
landscape around the Wabamun plant is mainly wooded 
with small marshes and an extensive area (5-10 km^) of 
excavated and reclaimed terrain to the north. The lake 
receives light recreational use, and some shoreline sec- 
tions are developed with cottages and small marinas. 

The nest box is in plain view from a public road that 
runs between the lake and the plant. Depending on 
weather conditions, we watched from different vantage 
points 0.1-2 km away, and we used 8X wide-angle bin- 
oculars and 20— 60X telescopes. Observation periods, last- 
ing 3-6 hr each, usually between mid-afternoon and sun- 
down, were spaced arbitrarily and increased during the 
period when the fledglings were on the wing. Over 7 yr, 
1998—2004, the number of observation days was 196; 15 
in June, 58 in July, 101 in August, 21 in September, and 
1 in October. The only two days of observation in 1997 
were added to the 1998 total. The median date when the 
first juvenile (always a male) fledged was 16 July (13-25 
July). The number of young fledged was three or four 
per season with a mean of 3.6 (N = 11 males and 14 
females). As indicated by their leg bands and plumage 
characteristics, the breeding pair consisted of the same 
individuals for the duration of the study, and we classified 
the female as a 2-yr old falcon in 1998 based on her 
brown dorsal color in 1997. The subspecific origin of 
these falcons is F. p. anatum and both originated from 
captive-reared stock nesting in the cities of Edmonton 


396 


Dekker and Taylor 


VoL. 39, No. 4 


and Calgary (Gordon Court, Alberta Environment, pers. 
comm.). 

Prey species upon which peregrines are known to feed, 
such as waterbirds and small passerines, are common in 
the study area. With a mass of 220-335 g (Dunning 
1984), the Franklin’s Gull {Larus pipixcan) is close to the 
mass limit that male peregrines can or are willing to carry 
over long distances. Although Franklin’s Gulls are locally 
scarce in early summer, large migrating flocks begin ar- 
riving in mid-July. Rock Pigeons (Columba livid), which 
are the dominant prey for peregrines in human altered 
environments (Ratcliffe 1993), were resident at the plant 
site but only in small numbers (<20). 

The terms hunt and attack are used interchangeably 
and represent one attempt at capturing prey including 
one or more stoops or passes at the same target of which 
the outcome was known. Tandem hunts were simulta- 
neous attacks by both adults on a flock of prey or an 
individual prey. Group hunts by a combination of parents 
and juveniles were tallied as adult hunts, and their kills 
were considered to have been made by the adults, al- 
though in a few cases it was the juvenile which actually 
seized (or was allowed to seize) the prey. Hunts by juve- 
niles alone were not tallied. 

Details of hunts and kills were entered into field diaries 
and annotated tabulations. Observer bias in comparing 
results between years, we believe, was not a factor as D. 
Dekker recorded observations over the 7-yr study period. 
R Taylor was associate observer the last 3 yr of the proj- 
ect. Data sets were compared for significance by chi- 
square Test of Independence with a Williams’ correction 
as described in Sokal and Rohlf (1981). 

Results 

Hunting Methods and Prey Species. We ob- 
served 386 hunts by the adult peregrines. Nearly 
all (99.5%) were directed at airborne prey and 
consisted of two primary methods: attacks 

launched from soaring flight (76.7%) or from a 
perch (23.3%; Table 1). The falcons typically be- 
gan a soaring sequence by circling up over the 
plant and using the hot air of the three stacks to 
gain height rapidly. Drifting downwind, the soaring 
falcons reached altitudes estimated to exceed 1000 
m. By flapping their wings or gliding, they cruised 
upwind. Attacks on prey flying lower than the fal- 
con were made by deep stoops with wings partly or 
fully flexed. Some of these attacks began with a 
burst of wing flaps and ended in a stoop, which 
could either be near-vertical or oblique. If the prey 
evaded the initial stoop, the falcon might follow up 
with one or more additional stoops or passes. 

Still-hunting attacks were launched from high 
perches such as the catwalk railing near the top of 
the 91 m stacks. With rapid wing flaps, these fal- 
cons headed for targets some distance away (>100 
m) . After unsuccessful or aborted attacks, the per- 
egrines commonly returned to the plant, either to 


perch or to regain altitude by soaring. Some hunt- 
ing sequences lasted 3-4 hr before a prey was cap- 
tured. On other days, we observed falcons catch 
three prey in less than 1 hr. 

The success rate of all hunts {N — 386) by the 
adult peregrines, either attacking prey solo or in 
tandem, was 30.3% (Table 1). Perch hunts were 
not significantly (G = 0.77, P = 0.379) more suc- 
cessful than soar hunts (35.5% and 28.7%, respec- 
tively). The success rate of the female was 37.1%, 
not significantly (G = 2.37, P — 0.124) different 
from that of the male (24.1%), but there was a 
notable difference in the taxa of the prey taken by 
the two sexes (Table 2). The male caught 84.4% 
of 32 small passerines, but only 22.9% of 62 gulls. 
Nearly all gulls seized by the male were released to 
his mate or a fledgling <1 sec after capture. If nei- 
ther the female nor any of the fledglings were 
nearby, the male, upon seizing a gull, brought it 
down steeply and left it on open ground or on the 
roof of the plant ( A = 5) . The female carried gulls 
with apparent ease over distances exceeding 1 km. 

The majority (84.9%) of gull kills observed at 
close range or found as prey remains (A = 73) 
were juveniles. Juvenile Franklin’s Gulls were often 
seized in mid air during the falcon’s first stoop. By 
contrast, adult gulls typically evaded a peregrine by 
rising steeply. Some evaded two or more additional 
passes and were eventually left alone. Others de- 
scended and plunged into water. In seven instanc- 
es, the female peregrine repeatedly swooped at the 
swimming gull. Three were retrieved from the wa- 
ter and carried off. Some white birds, assumed to 
be gulls, that evaded 20 or more diving attacks far 
out over the lake may have been Common Terns 
{Sterna hirundo). Two terns were carried to the 
plant, but their capture had not been observed. 
Ring-billed Gulls {Larus delawarensis) were some- 
times forced down, but we found no evidence that 
any were killed. 

Rock Pigeons were seldom attacked. On two oc- 
casions, the female stooped unsuccessfully at free- 
flying flocks of pigeons. Both adults made oppor- 
tunistic passes at pigeons that flushed from plant 
ledges below them, but the only successful pigeon 
hunts (A = 2) were initiated by fledglings. In at 
least one of these hunts, the capture was made by 
the adult female, who joined the attack. 

During the sixth year of study, we gained the im- 
pression that the falcons had become more suc- 
cessful — in particular, at capturing gulls — than dur- 
ing the preceding years. This impression was 


December 2005 


Foraging Success of Peregrine Falcons 


397 



•D 

10 


1998 1999 2000 2001 2002 2003 2004 

Years 

Figure 1. Hunting success per year of the same pair of 
Peregrine Falcons breeding in central Alberta, 1998— 
2004. The respective number of hunts/kills for the 7 yr 
of study were as follows: 108/23, 47/11, 25/8, 37/12, 44/ 
16, 61/22, 64/25. 

strengthened in year 7 and confirmed by subse- 
quent data analysis. The overall success rate of the 
adults increased from 21.3% in the first year of 
study to 39.1% in the seventh year of study (Fig. 
1 ). 

Tandem Hunts. Each year from mid June on- 
wards, both parents often began soaring together. 
The male gained height quicker and always circled 
higher than the female. Stooping alternately, they 
attacked 68 prey from soaring flights (Table 1). If 
no prey had been sighted for some time, the pair 
separated or headed off into the distance, one fol- 
lowing the other, either in soaring or flapping 
flight. Tandem attacks were also launched from 
high perches on the plant. In tandem attacks, both 
falcons could take the lead, and the male flew 
higher than the female. While the female ap- 
proached the prey directly, the male typically at- 
tacked from above in a near vertical stoop. A high 
percentage (60-75%) of tandem hunts were lost 
from view before we could either see the target or 
the result. Tandem hunts of which the outcome 


was known (N = 100) represented 25.9% of all 
hunts, and their success rate was higher than in 
286 solo hunts (39.0% versus 27.3%), but the dif- 
ference was not significant (G — 2.39, P = 0.121). 
The success rate of 32 tandem attacks launched 
from a perch was 50.0%, compared to 33.8% for 
tandem attacks from soaring position. 

Tandem hunts resulted in 39 kills (Table 1). Of 
these, 26 captures of prey were observed in detail: 
seven were seized by the female, 19 by the male. 
Prey caught by the male in tandem attacks were 
surrendered at once to the female, either by feet- 
to-feet transfer (N = 5) or released in the air (N 
= 14). Thirteen of these aerial drops were gulls, 
which resumed flight upon release. Nine were sub- 
sequently seized by the female; four evaded her 
passes and were let go or escaped by splashing 
down into water. 

Hunts by Groups of Adults and Fledglings. In 

23.3% of hunts (N = 386), one or both parents 
were accompanied, and often harassed, by one or 
more juveniles (Table 3) . Some fledglings closely 
followed the adults prior to the start of a hunt, 
others approached hurriedly from a distance to 
join a hunt in progress. A significant majority of 
group hunts (G = 37.96, P = 0.001) were led by 
the adult male, which made 71.0% of the 31 kills 
(Table 3). Prey caught by the parents were trans- 
ferred, usually at once, to the approaching juve- 
niles, either by feet-to-feet transfers (N = 15) or 
through aerial drops (N = 7) . At least three small 
passerines that were dropped were still alive and 
resumed flight. They were recaptured by a juvenile 
or by the adult and released again. Three live- 
drops were gulls. In addition, four of 11 gulls 
caught in adult-juvenile hunts were seized in flight 
by the juveniles “chaperoned” by adults. 

After successful group hunts, one or more juve- 
niles fed on the kill. If the feeding site was in an 
open area away from the plant, the adult female 
typically perched on a pole nearby {N = 14). At 


Table 1. Hunts and kills made by a pair of Peregrine Falcons nesting on a power plant in central Alberta, 1998- 
2004. The hunting success rates are presented in parentheses. 



Hunts/Kills from Soaring 

Hunts/Kills from Perch 

Total Hunts 

Adult male (24.1%) 

176/44 (25.0%) 

40/8 (20.0%) 

216/52 

Adult female (37.1%) 

52/18 (34.6%) 

18/8 (44.4%) 

70/26 

In tandem hunts (39.0%) 

68/23 (33.8%) 

32/16 (50.0%) 

100/39 

Totals (30.3%) 

296/85 (28.7%) 

90/32 (35.5%) 

386/117 


398 


Dekker and Taylor 


VoL. 39, No. 4 


the approach of people, she vocalized and 
“swooped” overhead. She also stooped at crows or 
Buteo hawks, apparently to drive them away. After 
the fledglings were satiated and left, the adult fe- 
male sometimes fed on the remains of the kill {N 
= 5). 

Discussion 

Hunting Success. The hunting success of adult 
peregrines on breeding territory is generally high- 
er than that of migrating or wintering peregrines 
(Dekker 1980, Roalkvam 1985). Jenkins (2000) 
compared data presented in eight publications on 
breeding peregrines and found an extreme varia- 
tion (9.3-84.1%) in hunting success, but much less 
variation in hunting methods. The majority (58- 
75%) of attacks summarized in his review were 
launched from a perch. By contrast, the percent- 
age of perch hunts was only 23% at Wabamun. The 
“still-hunting strategy” of perch hunts reduces the 
energy cost of foraging. However, high-soaring 
flight is also relatively energy-efficient and widens 
the radius of the attack zone (Enderson and Craig 
1997). Soaring flight has been reported as a com- 
mon hunting method in many regions, but not to 
such a high degree as documented in this study. 
The second highest use was recorded in Africa, 
where breeding pairs launched 30-42% of hunts 
from flight (Thiollay 1988, Jenkins 2000). The use 
of factory exhaust to facilitate soaring flight by the 
Wabamun peregrines was described in an earlier 
publication (Dekker 1999), but this phenomenon 
has to our knowledge not been reported else- 
where. 

It is noteworthy that the falcons in this study at- 
tacked nearly all of their prey in flight, often at 
great altitudes. This is in sharp contrast to migrat- 
ing or wintering peregrines, which commonly use 
surprise methods to attack prey on the ground or 
in shallow water (Dekker 1980, 2003, Cresswell 
1996). However, there is also an element of sur- 
prise involved if a soaring peregrine stoops from a 
great height at prey flying far below. In solo hunts, 
the male frequently made long stoops that levelled 
out low over woods and were aimed at flocks of 
small passerines flying just beyond the trees. Sur- 
prise is probably also a factor in tandem hunts in 
which the male stoops from high above while the 
female pursues the prey at a lower altitude. 

A key factor in the hunting success of peregrines 
is the vulnerability of individual prey, which is dif- 
ficult to assess for the human observer. It is well- 


known that predation risk for land birds increases 
over water (Herbert and Herbert 1965). Converse- 
ly, water birds become vulnerable over land (Hunt 
et al. 1975, Dekker 1980). Mature prey on home 
territory should be harder to catch for a raptor 
than juvenile prey passing over unfamiliar terrain. 
In central Alberta, the migrations of juvenile prey 
species approximately coincide with the period 
when fledgling peregrines are on the wing and 
when the parental task of provisioning them is 
most demanding. In this study, the peregrines were 
very successful at capturing juvenile Franklin’s 
Gulls. For instance, on 7 d between 17 July and 9 
August 2003, when thousands of gulls were passing 
through the area, the adult female, hunting solo, 
captured each of seven juvenile gulls in her first 
attack of the afternoon and each requiring only 
one stoop. While it does not seem surprising that 
bird-hunting falcons should become better at what 
they do as they become older and gain in experi- 
ence, data in support of that notion have, to our 
knowledge, not been published before. An expla- 
nation for the year-to-year increase in the success 
rate of the Wabamun pair is that these falcons be- 
came specialists on juvenile Franklin’s Gulls, which 
differ from adults by their dusky color and absence 
of black on the head. Juvenile targets were proba- 
bly pre-selected before the falcons began their at- 
tack. A similar pre-selection hypothesis was ad- 
vanced for the high success rate (73-93%) of “Red 
Baron,” an adult male peregrine hunting over 
coastal marshes in New Jersey (Cade and Burnham 
2003:333). 

In this study, the prey component changed sig- 
nificantly after the first year (Tables 4 and 5). In 
the first year of study, gulls made up only 8.7% of 
kills compared to 63.8% in years 2-7 (G = 10.85, 
P = 0.001). The proportion of small passerines 
changed significantly from 60.9% in the first year 
to 19.1% in years 2-7 (G = 7.01, P — 0.008). Over- 
all, the capture success on gulls (of both age 
groups) was significantly higher (G = 5.35, P = 
0.021) than on small passerines (42.5% versus 
24.1%). Successful taking of juvenile gulls may be 
even higher than for adults, but an adequate sam- 
ple was not available for adults. As reported in re- 
sults, the sample of gull kills {N = 73) included 
84.9% juveniles. 

Coincident with the observed prey switch, we re- 
corded a major change in foraging participation 
between the sexes after the first year (Table 6). 
The male’s solo hunts in the first year were 78.7% 


December 2005 


Foraging Sugcess of Peregrine Falgons 


399 


Table 2. Prey taxa captured by the Peregrine Falcon pair hunting solo or in tandem. 



Franklin’s 

Gulls 

Small 

Passerines 

Smalt. 

Shorebirds 

Ducks 

Other or 
Unidentified 

Adult male 

14 

27 

1 

1 

9 

Adult female 

21 

2 

1 

2 

0 

Pair in tandem 

27 

3 

5 

0 

4 

Totals 

62 

32 

7 

3 

13 


of total hunts (N = 108), significantly greater (G 
= 8.08, P = 0.004) than the 47.1% of hunts in 
years 2-7 {N = 278). By contrast, the female’s 
share during the first year was 2-7%, and this rose 
significantly to 24.1% in years 2-7 (G = 23.51, P = 
0.001). A possible explanation is that the female 
lacked skill and experience in her first year on ter- 
ritory. As she became older and gained expertise, 
she began to play a more active role in foraging. 
There was no significant change (G = 2.70, P = 
0.1) in the proportion of tandem hunts between 
the first year and years 2-7 (Table 6) . 

Our findings that the primary prey of the male 
was passerines, whereas the female’s main prey was 
gulls (Table 2), lends support to the hypothesis 
that the reversed sexual size dimorphism in raptors 
such as seen in the peregrine allows them to ex- 
ploit a wider prey base and reduces competition 
between the sexes (Selander 1966). 

Cooperative Hunting. Are peregrines that attack 
the same prey simultaneously with their mates or 
fledglings actually cooperating? Or, are the individ- 
uals simply reacting at the same time to the stim- 
ulus of sighting prey? True cooperative hunting 
differs fundamentally from other forms of group 
predation, such as pseudo-cooperative hunting (El- 
lis et al. 1993). In true cooperative hunting, the 
group consists of at least two members that are a 
stable social unit, and their cooperation should 
benefit the group rather than just the individual. 
An example of true cooperative foraging is de- 
scribed by Bednarz (1988) for the Harris’s Hawk 


{Parabuteo unicinctus). Ellis et al. (1993) conclude 
that cooperative hunting is the most efficient strat- 
egy for capturing prey in many situations and that 
each form of social foraging should have evolved 
as an adaptive advantage enhancing the fitness of 
all individuals in the group. In the Aplomado Fal- 
con (Falco femoralis), pair hunting is more than 
twice as productive as solo hunting when the prey 
are birds (Hector 1986). Similarly, the success rate 
of pair hunting in peregrines may depend on the 
species of prey hunted (Jenkins 2000). 

In this study, tandem hunts by the mated pair of 
peregrines were more successful than solo hunts 
(39.0 versus 27.3%), although the difference was 
not statistically significant. Moreover, the individu- 
al success rate of the two partners in tandem hunts 
is actually halved. So what, if any, is the fitness value 
of tandem hunting for this pair of breeding pere- 
grines? It may lie in the fact that any action by the 
male that benefits his mate or their progeny 
should be of adaptive advantage. 

At Wabamun, the adult male was not seen to eat 
the gulls he killed, although on two occasions he 
fed on gull remains abandoned by fledglings. The 
fact that he mainly hunted such prey in the com- 
pany of his mate or fledglings suggests that his role 
was a cooperative one. This view is supported by 
anecdotes such as the following. By mid September 
1999, after the fledglings had dispersed, the adults 
were still hunting together. One evening, shortly 
after both had soared up over the plant, the male 
caught a small passerine and landed on a pylon to 


Table 3. Hunts/kills made by one or both Peregrine Falcon parents accompanied by one or more fledglingjuveniles. 



1 Juvenile 

2 Juveniles 

3 Juveniles 

Totals 

Adult Male 

29/10 

34/10 

6/2 

69/22 

Adult female 

3/2 

6/4 

0/0 

9/6 

Both adults 

6/3 

5/0 

1/0 

12/3 

Totals 

38/15 

45/14 

7/2 

90/31 


400 


Dekker and Taylor 


VoL. 39, No. 4 


Table 4. Prey captured by a pair of Peregrine Falcons as a percentage of total prey per year. 


Year 

Percent Gulls 

Percent Passerines 

Percent Other 

No. Total Prey 

No. Hunts 

1 

8.7 

60.9 

30.4 

23 

108 

2 

63.6 

27.3 

9.1 

11 

47 

3 

62.5 

12.5 

25.0 

8 

25 

4 

58.3 

16.7 

25.0 

12 

37 

5 

43.7 

31.3 

25.0 

16 

44 

6 

63.6 

27.3 

9.1 

22 

61 

7 

80.0 

4.0 

16.0 

25 

64 

Totals 

53.0 

27.4 

19.6 

117 

386 


consume his prey. Instead of interfering, the fe- 
male perched on the next pylon and waited for her 
mate to finish his meal. Presently, both soared up 
again and eventually flew out high over the lake, 
stooping in tandem at gulls. Like many similar ob- 
servations, this incident suggests that the male’s 
role was truly a cooperative one — to assist his mate 
in her foraging. 

The question of whether attacks on the same 
prey by a combination of adults and fledglings can 
be considered cooperative seems even less clear, 
for it is apparent that the juvenile is primarily in- 
tent on kleptoparasitizing the adult. If there is co- 
operation, it is one-sided and benefits only the ju- 
veniles (i.e., this may be considered as parental 
investment). Nevertheless, the evolutionary value 
of adult/ fledgling combinations is undeniable, as 
a well-fed progeny enhances the future of the fam- 
ily genes. In support of the hypothesis that the par- 
ticipation of adults in joined hunts with their fledg- 
lings indeed represents parental care, we present 
a number of anecdotal observations in the appen- 
dix. 

Live-drops of Prey. Cade (1982) and Sherrod 
(1983) characterize the behavior of paired male 
peregrines in social interactions with their bigger 
mate as submissive and even fearful. Females take 
food from males in an aggressive manner, not only 


Table 5. Comparison of main prey taxa captured by a 
pair of Peregrine Falcons in study years 1 and 2-7. 



Prey Taxa 

No. 

Kills 

Percent of 
Totals Kills 

N 

Year 1 

Gulls 

2 

8.7 

23 

Years 2-7 


60 

63.8 

94 

Year 1 

Passerines 

14 

60.9 

23 

Years 2-7 


18 

19.1 

94 


during the breeding season but also on migration 
or at the wintering grounds. Unmated females rou- 
tinely force males to drop just-caught prey (Dekker 
1980, 2003). Tandem hunting by unmated pere- 
grines is relatively common and driven by compe- 
tition rather than cooperation. Two or more (up 
to six) migrating or wintering peregrines were seen 
in joint pursuit of the same prey in Alberta and 
British Columbia. Male peregrines were forced to 
make live-drops of prey not only to female conspe- 
cifics, but also to other kleptoparasitic raptors, 
such as Prairie Falcons {Falco mexicanus) and Bald 
Eagles {Haliaeetus leucocephalus; Dekker 1995, 
1998). 

I concur with Sherrod (1983) that the release of 
just-caught prey by the male peregrine at the ap- 
proach of an aggressive female may be triggered 
by the impulse to avoid close contact. The impor- 
tance of a timely release was demonstrated on 21 
August 2003, when the Wabamun male, carrying a 
just-caught prey and flying 20-25 m high, was met 
by a screaming juvenile female in typical begging 
flight. Apparently, the male was just too late in re- 
leasing his prey, for he was seized by the juvenile. 
Locked by their feet, both birds tumbled about 15 


Table 6. Number of hunts by the adult male and fe- 
male, either hunting solo or in tandem, in year 1 and 
years 2-7. 




No. 

Hunts 

Percent 
OF Total 

Year 1 

Adult male 

85 

78.7 

N = 108 hunts 

Adult female 

3 

2.7 


Tandem 

20 

18.5 

Years 2-7 

Adult male 

131 

47.1 

N = 278 hunts 

Adult female 

67 

24.1 


Tandem 

80 

28.8 


December 2005 


Foraging Success of Peregrine Falcons 


401 


m before separating again, while the prey item fell 
into the bushes below and was lost. 

Based on a thorough review of the literature and 
on his own studies, Sherrod (1983) observed that 
the behaviors of parents and fledglings comple- 
ment each other. While the youngster only wants 
food, the adult appears to be very willing to supply 
that food. Some live-drops are no doubt the result 
of the aggressive approach of the begging falcon. 
However, as argued by Newton (1979), other live- 
drops are clearly intentional as parent peregrines 
will recapture the live-dropped prey and release it 
again if the fledgling fails to catch it the first time. 
Such repeat drops were seen in this and other stud- 
ies (Frank 1994, Treleaven 1998). 

Conclusion 

Some of the aerial drops of just-caught and still 
live prey from adult to fledgling (described in the 
appendix) lend support to the hypothesis that the 
parents were teaching their young by example. Al- 
though young falcons do not need to be taught to 
hunt, such extended parental care might well give 
them a certain survival advantage after leaving the 
nest site. The observation that live-drops were also 
made from adult to adult, both at Wabamun and 
at a natural nest site in northern Alberta (Dekker 
1999), may seem to refute the above hypothesis. 
Because there can be no doubt that adult females 
are accomplished hunters, the male’s reason for 
making live-drops to his mate can have nothing to 
do with teaching. However, here a secondary and 
complementary factor comes into play. Whether 
hunting with his mate or a fledgling, the male plays 
a dual role, either in support of his mate or prog- 
eny. 

It is difficult, if not impossible, to translate all of 
the anecdotes detailed in the appendix into hard 
data in support of the teaching hypothesis. As to 
the alternative, but not exclusive, hypothesis that 
parent peregrines assist their fledglings in captur- 
ing their first prey, on the evidence presented 
here, we are convinced that the answer is affirma- 
tive. 

ACKNOWI.EDGMENTS 

I thank T.J. Cade for his review of an earlier draft of 
this paper. Referees G. Malar and I. Newton made very 
helpful comments on the second version. Throughout, 
the input ofj. Bednarz was particularly valuable. W. Nel- 
son provided relevant information on nest box installa- 
tion and the weight of Franklin’s Gulls. G. Court did the 
statistical tests. R. Dekker drafted the figure. Occasional 


co-observers were G. Court, I. Dekker, K. Kunst, and J 
Morrison. I thank TransAlta Utilities for access to obser- 
vation points out of bounds to the general public. All 
expenses incurred during this study were privately fund- 
ed by the authors except during the last year, when D 
Dekker received travel grants from TransAlta Utilities 
and the Alberta Conservation Association. 

Literature Cited 

Bednarz, J.C. 1988. Cooperative hunting in Harris’s 
Hawks {Parabuteo unicinctus) . Science 2^9:1525— 1527 . 
Bird, D.M. and Y. Aubry. 1982. Reproductive and hunt- 
ing behavior in Peregrine Falcons in southern Que- 
bec. Can. Field-Nat. 96:167-171. 

Brown, L. and D. Amadon. 1968. Eagles, hawks and fal- 
cons of the world. Country Life Books. Hamlyn 
House, Feltham, England. 

Cade, TJ. 1960. Ecology of the peregrine and Gyrfalcon 
populations in Alaska. Univ. Calif. Publ. Zool. 63:1561- 
290. 

. 1982. The falcons of the world. Cornell Univer- 
sity Press, Ithaca, NY U.S.A. 

and W. Burnham. 2003. Return of the peregrine. 

The Peregrine Eund, Boise, ID U.S.A. 

Cresswell, W. 1996. Surprise as a winter hunting strategy 
in Sparrowhawks, peregrines, and Merlins. Ibis 138. 
684-692. 

Dekker, D, 1967. Disappearance of the Peregrine Falcon 
as a breeding bird in a river valley in Alberta. Blue fay 
30:175-176. 

. 1980. Hunting success rates, foraging habits and 

prey selection of Peregrine Falcons migrating through 
central Alberta. Can. Field-Nat. 94:371-382. 

. 1995. Prey capture by Peregrine Falcons winter- 
ing on southern Vancouver Island, British Columbia 
J. Raptor Res. 29:26-29. 

. 1998. Over-ocean flocking by Dunlins and the 

effect of raptor predation at Boundary Bay, British 
Columbia. Can. Field-Nat. 112:694—697. 

. 1999. Bolt from the blue. Wild peregrines on the 

hunt. Hancock House Publishers. Surrey, BC Canada, 
Blaine, WA U.S.A. 

. 2003. Peregrine Falcon predation on Dunlins 

and ducks and kleptoparasitic interference from Bald 
Eagles wintering at Boundary Bay, British Columbia. 
J. Raptor Res. 37:91-97. 

Dunning, J.B. 1984. Body weights of 686 species of North 
American birds. Western Bird Banding Assoc. Mo- 
nogr. No 1. 

Ellis, D.H., J.C. Bednarz, D.G. Smith, and S.P. Flem- 
ming. 1993. Social foraging classes in raptorial birds 
BioScience 43:14—20. 

Enderson, J.H. and G.R. Craig. 1997. Wide-ranging by 
nesting Peregrine Falcons determined by radio-telem- 
etry. /. Raptor Res. 31:333-338. 

Frank, S. 1994. City peregrines. A ten-year saga of New 
York city falcons. Hancock House Publishers. Surrey, 
BC Canada, Blaine, WA U.S.A. 


402 


Dekker and Taylor 


VoL. 39, No. 4 


Fyfe, R.W. 1976. Rationale and success of the Canadian 
Wildlife Service peregrine breeding project. Can. 
Field-Nat. 90:308-319. 

Hector, D.P. 1986. Cooperative hunting and its relation- 
ship to foraging success and prey size in an avian 
predator. Ethology 73:247-257. 

Herbert, R.A. and K. Herbert. 1965. Behavior of Pere- 
grine Falcons in the New York City region. Auk 82: 
62-94. 

Holroyd, G.L. and U. Banasch, 1990. The reintroduc- 
tion of the Peregrine Falcon into southern Canada. 
Can. Field-Nat. 104:203-208. 

Hunt, W.G., R.R. Rogers, and D.J. Slowe. 1975. Migra- 
tions and foraging habits of Peregrine Falcons on the 
Texas coast. Can. Field-Nat. 89:11—123. 

Jenkins, A.R. 2000. Hunting mode and success of African 
peregrines: does nesting habitat quality affect forag- 
ing efficiency? Ibis 142:235-246. 

Newton, I. 1979. Population ecology of raptors. Buteo 
Books, Vermillion, SD U.S.A. 

Palmer, R.S. 1988. Handbook of North American birds. 
Vol. 5. Diurnal raptors, part 2. Yale University Press, 
New Haven, CT U.S.A. 

Ratclifee, D. 1993. The Peregrine Falcon. Poyser Pub- 
lishing. London, England. 

Roalkvam, R. 1985. How effective are hunting pere- 
grines? Raptor Res. 19:27-29. 

Selander, R.K. 1966. Sexual dimorphism and differential 
niche utilization in birds. Coracior 68:113-151. 
Sherrod, S. 1983. Post-fledgling behavior of the Pere- 
grine Falcon. The Peregrine Fund, Ithaca, NY U.S.A. 
SoKAL, R.R. and F.J. Rohlf. 1981. Biometry: the princi- 
ples and practice of statistics in biological research. 
Freeman and Company, San Francisco, CA U.S.A. 
Thioliay, J.-M. 1988. Prey availability limiting an island 
population of Peregrine Falcons in Tunisia. Pages 
701-710 m TJ. Cade, J.H. Enderson, C.G Thelander, 
and C.M. White [Eds.], Peregrine Falcon populations: 
their management and recovery. The Peregrine Fund, 
Boise, ID U.S.A. 

Treleaven, R.B. 1998. In pursuit of the peregrine. Tier- 
cel Publishing, Wheathampstead Herts, England. 
White, C.M, NJ. Clum, TJ. Cade, and W.G. Hunt. 2002. 
Peregrine Falcon {Falco peregrinus). In The birds of 
North America, No. 660. A. Poole and F. Gill [Eds.], 
The Birds of North America, Inc., Philadelphia, PA 
U.S.A. 

Received 18 September 2003; accepted 15 July 2005 

Appendix. Observations of parental care by a pair of per- 
egrines toward their fledglings and nearly-fledged juve- 
niles. 

(1) On 16 August 1999, having stooped at a gull and 
missed his target, the adult male maneuvered under it in 
a peculiarly slow flying style. He appeared to drive the 
gull upward and prevent it from dropping into the water 


below. Seconds later, a juvenile female stooped from 
above, just missing the gull. The adult male then stooped 
again, just short of the target. Chased by both peregrines, 
the gull eventually escaped. In this case, as in others, it 
seemed clear that the male did not want to catch the gull 
himself. Instead, he attempted to set it up for the ap- 
proaching juvenile. 

(2) On two occasions, when a gull was chased by a 
group of one adult and one fledgling, the adult hit or 
touched the prey, causing it to cartwheel in the air. In 
the next instant, the gull was seized by the closely-follow- 
ing juvenile. Our impression was that the adults were not 
intent on taking the gull themselves. 

(3) On 12 August 2000, the adult male, while soaring 
in tandem with his mate, suddenly sprinted away (400— 
500 m) to hit a shorebird that was being chased by two 
juveniles at an altitude estimated at 75 m. Dead or 
stunned, the prey fell into reeds and was lost before the 
juveniles could recover it. This was one of only four in- 
stances when these peregrines hit their prey in the air, 
as opposed to the common method of binding to the 
prey and carrying it down. The remarkable point is that 
the adult male did not stoop and recover the plunging 
shorebird himself, which would have been a normal pro- 
cedure if he had been hunting alone. 

(4) A Rock Pigeon pursued by a fledgling was taken 
by the adult female, who transferred it at once to the 
juvenile. On their own, the adults seldom pursued pi- 
geons. 

(5) Fledglings that have been on the wing for only a 
few days have difficulty carrying a gull. After making an 
aerial transfer to an unskilled juvenile, the adults will ac- 
company it in flight and retrieve the gull if dropped. By 
the same token, seemingly aware of the lack of ability in 
a badgering female youngster, the adult female refused 
to transfer a just-caught gull and instead carried it ca. 
500 m to the roof of the plant, where she at once sur- 
rendered it to the fledgling. 

(6) The adult male demonstrated an awareness of the 
different needs of the genders. In group hunts with male 
offspring, he selected mostly passerines or small shore- 
birds (41 of 45). By contrast, he usually attacked gulls in 
the company of females. After capturing one or more 
gulls for his mate or the fledglings, he hunted smaller 
prey for himself. 

The above interactions between parents and fledglings 
parallel numerous anecdotes described by long-term per- 
egrine observers in New York and Great Britain (Frank 
1994, Treleaven 1998). 

It is our impression that adult peregrines appear to 
know how to assist their progeny in gaining indepen- 
dence. For a start, there is a change in food exchanges 
when the first of the juveniles are close to fledging. Both 
at Wabamun and a natural nest site in central Alberta, 
the adult male appeared to coax a fully-feathered male 
youngster to fly. Instead of transferring the prey item in 
the usual direct manner, he cruised back and forth just 


December 2005 


Foraging Success of Peregrine Falcons 


403 


out of reach, holding the prey in his lowered feet. His 
behavior cannot be explained as having been prompted 
by fear of a bigger bird for this incident involved a male 
juvenile. After the young male failed to fledge, the adult 
gave the food item to another juvenile, a fully-feathered 
female sitting on the catwalk. Similarly, on two other oc- 
casions when he arrived with a small prey, he flew right 
up to a screaming juvenile male that was perched on the 
edge of a catwalk, but instead of exchanging the prey 
feet-to-feet, he dropped it and recovered the falling item 
in a quick stoop. He repeated this teasing show three or 
four times. After the fully-feathered juvenile male failed 


to fly, the adult gave the food item directly to a less-de- 
veloped female still in the nest box. 

Another noteworthy exchange involved the female 
One day, after she had delivered a prey to two juvenile 
males, which shared the food, the male arrived, also with 
prey. He gave it directly to one of the same two young- 
sters. Instantly, the adult female interfered. She took the 
food item away from the young male and offered it in- 
stead to a juvenile female which had not been fed for 
several hours. Similar incidents were seen at Wabamun 
and at a natural nest site in central Alberta (D. Dekker 
unpubl. data). 


J Raptor Res. 39(4):404-416 
© 2005 The Raptor Research Foundation, Inc. 

NESTING ECOLOGY AND BEHAVIOR OF BROAD-WINGED HAWKS 
IN MOIST KARST FORESTS OF PUERTO RICO 

Derek W. Hengstenberg and Francisco J. VilellaI 

USGS Biological Resources Division, Cooperative Research Units, MS 9691, Department of Wildlife and Fisheries, 

Mississippi State University, Mississippi State, MS 39762 US. A. 

Abstract. The Puerto Rican Broad-winged Hawk {Buteo platypterus brunnescens) is an endemic and 

endangered subspecies inhabiting upland montane forests of Puerto Rico. The reproductive ecology, 
behavior, and nesting habitat of the Broad-winged Hawk were studied in Rio Ab^o Forest, Puerto Rico, 
from 2001-02. We observed 158 courtship displays by Broad-winged Hawks. Also, we recorded 25 ter- 
ritorial interactions between resident Broad-winged Hawks and intruding Red-tailed Hawks {Buteo ja- 
maicensis jamaicensis) . Broad-winged Hawks displaced intruding Red-tailed Hawks from occupied terri- 
tories (P = 0.009). Mayfield nest survival was 0.67 across breeding seasons (0.81 in 2001, N = 6; 0.51 
in 2002, N = 4), and pairs averaged 1.1 young per nest (years combined). The birds nested in mixed 
species timber plantations and mature secondary forest. Nests were placed in the upper reaches of large 
trees emerging from the canopy. Nest tree DBH, understory stem density, and distance to karst cliff wall 
correctly classified (77.8%) nest sites. 

Key Words: Broad-wing Hawk; Buteo platypterus brunnescens; endangered; nest success; prey delivery; hab- 

itat model; karst forest; Puerto Rico. 


ECOLOGIA REPRODUCTIVA DE BUTEO PiJiTYPTERUS BRUNNESCENS EN BOSQUES DE CALIZA 
HUMEDOS DE PUERTO RICO 

Resumen. Buteo platypterus brunnescens es una subespecie de rapaz endemica a los bosques montanos 

de Puerto Rico. Investigamos la ecologia reproductiva y el comportamiento de B. p. brunnescens en el 
Bosque de Rio Abajo, Puerto Rico, durante 2001 y 2002. Observamos 158 despliegues de cortejo en 
Rio Abajo. Observamos 25 encuentros territoriales entre B. p. brunnescens y B. jamaicensis jamaicensis. B. 
p. brunnescens desplazo a B. j. jamaicensis de sus territorios el 84% de las veces (P = 0.009). La supervi- 
vencia de los nidos en ambas temporadas fue de 0.67 (0.81 en 2001, N = 6; 0.51 en 2002, N= 4). Los 
nidos produjeron un promedio de 1.1 volantones por nido (afios combinados). Encontramos nidos en 
areas de bosque secundario maduro y plantaciones forestales. La altura del dosel, diametro del arbol, 
densidad del sotobosque y distancia a farallon de mogote clasificaron correctamente (77.8%) los nidos 
en Rio Abajo. 

[Traduccion de los autores] 


The Broad-winged Hawk {Buteo platypterus brun- 
nescens) is an endemic woodland raptor of upland 
montane forests of Puerto Rico. This subspecies is 
listed as endangered (Federal Register 1994) hy 
the Puerto Rico Department of Natural and Envi- 
ronmental Resources (DNER) and the U.S. Fish 
and Wildlife Service (USFWS). The Broad-winged 
Hawk in Puerto Rico is nonmigratory and exhibits 
a limited geographic range with known popula- 
tions restricted to montane forests (Delannoy 
1997). Breeding in Puerto Rico begins in late De- 
cember, with nests placed in the upper reaches, but 


^ Corresponding author email address; Mlella@cfr. 
msstate.edu 


below the high canopy (Delannoy and Tossas 
2002). This insular subspecies is smaller and dark- 
er than its North American nominate counterpart 
Buteo platypterus platypterus, but larger than the 
Lesser Antillean subspecies (Raffaele 1989, Good- 
rich et al. 1996). The most recent population es- 
timate for the Broad-winged Hawk in forest re- 
serves of Puerto Rico is approximately 125 
individuals (Delannoy 1997). 

The nesting biology of the Broad-winged Hawk 
in North America has been described by a number 
of authors (e.g., Fitch 1974, Matray 1974, Rosen- 
held 1984, Titus and Mosher 1987, Crocoll and 
Parker 1989). However, knowledge on the repro- 
ductive biology of the insular endemic subspecies 


404 


December 2005 


Puerto Rican Broad-winged Hawk Ecology 


405 



BWHA nest sites 


A 2001 

• 2002 



Forest trails 

Roads 

Rivers 

Lago Dos Bocas 
Rio Abajo Forest 


N 



900 0 900 Meters 



Figure 1. Locations of Broad-winged Hawk nest sites during the breeding season of 2001 and 2002 in Rio Abajo 
Forest, Puerto Rico. 


of Puerto Rico, and forest raptors of the West In- 
dies in general, are limited (Goodrich et al. 1996). 
Similarly, the available information on nesting be- 
havior and diet of Broad-winged Hawks in Puerto 
Rico is based on a few observations by Snyder and 
Kepler (1987). 

Additional information on the Puerto Rican 
Broad-winged Hawk’s reproductive ecology and 
nest habitat use is required to better understand 
current limiting factors, and provide recommen- 
dations for future research on habitat conservation 
in public and private lands. Herein, we report on 
the nesting ecology and behavior of the Broad- 
winged Hawk in a forest reserve in the moist lime- 
stone forest region of Puerto Rico. Specifically, we 
provide baseline information on courtship behav- 
ior and territorial defense, nest success and pro- 
ductivity, prey delivery rates by nesting pairs, and 
their habitat use. Moreover, we discuss the impli- 
cations our results on interactions between the 
Broad-winged Hawk and Red-tailed Hawk {Buteoja- 


maicensis jamaicensis) may have on future plans to 
establish a second wild population (by releasing 
captive-reared individuals) of the critically endan- 
gered Puerto Rican Parrot {Amazona vittata) in the 
Rio Abajo Forest (White et al. 2005). 

Study Area and Methods 

Study Area. We studied Broad-winged Hawks in the Rio 
Abajo Forest and surrounding private lands in Puerto 
Rico from 2000-02 (Fig. 1). The Rio Abajo Forest 
(18°20'N, 66°42'W) is managed by the Forestry Division 
of DNER and is in west-central Puerto Rico within the 
moist limestone region of the island (Ewel and Whitmore 
1973). This forest reserve comprises an area of 2300 ha 
with elevations ranging from 200-420 m. We obtained 
climate data for the study period from the site closest to 
our study area, the Dos Bocas NOAA weather station 
(NOAA 2002). 

Annual precipitation during our study averaged 18.3 
cm (range = 6.9-34.9 cm) in 2001 and 14.7 cm (range 
= 5.2-45.1 cm) in 2002. Mean annual temperature was 
25.3°C (range = 19.9— 30.6°C) in 2001 and 25.5°C (range 
= 20.1-30.9°C) in 2002. 

The rugged limestone region (i.e., karst) of Puerto 



406 


Hengstenberg and Vilella 


VoL. 39, No. 4 


Rico encompasses 27.5% of the island’s surface (Lugo et 
al. 2001). Topography in this region is extreme and char- 
acterized by karst formations of subterranean drainages, 
caves, dome shaped hills or “mogotes,” and deep sink- 
holes. Karst forest contains the largest tree species rich- 
ness of Puerto Rico (Lugo et al. 2001). Rio Abajo Forest 
IS fragmented on the eastern end by a double lane high- 
way and in the south-central part by a small community 
(Fig. 1) . About 75% of the forest is within the subtropical 
wet zone, the remaining quarter lies within the subtrop- 
ical moist zone (Ewel and Whitmore 1973). 

Previous studies indicated Broad-winged Hawks in 
Puerto Rico have a limited geographic range, and their 
abundance is higher in the Karst region compared to 
other life zones on the island (Delannoy 1997). As our 
primary objectives were to expand current knowledge on 
breeding ecology, habitat use, and movement patterns of 
Broad-winged Hawks (Hengstenberg and Vilella 2004), 
we selected the Rio Abajo Forest. Moreover, we were in- 
terested in comparing our results with findings of re- 
cently completed studies in Rio Abajo Forest (Delannoy 
and Tossas 2002). 

Vegetation of Rio Abajo is comprised of a mix of sec- 
ondary growth forests and timber plantations (Cardona 
et al. 1987). The midstory of some areas of secondary 
forest was characterized by abandoned shade-grown cof- 
fee and citrus plantations. Forest overstories were domi- 
nated by moca {Andira inermis) , capa prieto ( Cordia allio- 
dora), and guaraguao {Guarea guidonia). Approximately 
6.9% of the forest (171.7 ha) are managed timber plan- 
tations. Some stands are actively maintained (open un- 
derstory), while others had a developing understory. 
Timber plantations, approximately 30-50 yr old, of Hon- 
duras mahogany {Swietenia macrophylla) , maria {Calophy- 
llum brasiliense) , teca (Tectona grandis), and mahoe {Hibis- 
cus elatus) occur in valley bottoms and along mid-slopes 
(Cardona et al. 1987). 

As part of additional research efforts, we trapped 
Broad-winged Hawks in Rio Abajo Forest (Hengstenberg 
and Vilella 2004). Between January 2001 and July 2002, 
we trapped eight Broad-winged Hawks in the Rio Abajo 
Forest. We used octagonal and quonset style bal-chatri 
traps (Berger and Mueller 1959, Erickson and Hoppe 
1979), a modified dho-gaza trap (Hamerstrom 1963, 
Clark 1981) with a live Red-tailed Hawk, and a Rock Pi- 
geon {Columba livia) with noose harness. 

Each captured individual was banded with a unique 
color-coded leg band on the left leg and a standard U.S. 
Geological Survey (USGS) Bird Banding Laboratory 
band on the right leg. We also recorded morphometric 
measurements and determined gender. We collected a 
small amount (±5 mL) of blood from the brachia vein 
of the left wing of every captured individual and deter- 
mined gender through DNA typing. Each bird was fitted 
with a backpack mounted radio transmitter attached via 
a break-away backpack harness and a leather keel patch 
(Vekasy et al. 1996), Furthermore, we used visible size 
differences and markings to separate individual members 
of a broad-wing pair. 

Territorial Defense. We documented territorial en- 
counters between resident Broad-winged Hawks at Rio 
Abajo Forest and surrounding lands. Moreover, to ex- 
amine the behavioral interactions of the two sympatric 


Buteos, we observed territorial interactions between 
Broad-winged Hawks and intruding Red-tailed Hawks 
from limestone hilltops. We recorded the behavior of 
both birds (aggressive or passive) and the end result (de- 
terred or not deterred) . We used a binomial sign test for 
a single sample (Daniel 1990, Sheskin 2000) to test the 
hypothesis that both species of raptors displaced the oth- 
er randomly during aerial displays. 

Nest Searches and Monitoring. Broad-winged Hawk 
use areas were delineated through direct observations 
from limestone hills and hawk locations were plotted on 
USGS topographic quadrangles. We then extensively 
searched areas with radio-marked adults and document- 
ed aerial displays or other territorial behavior. Spot maps 
and historical nesting information were used to extend 
our searches into other potential nesting territories. All 
potential nests were monitored for reproductive activity 
beginning in February of each year. 

When a nest site was located, we built observation 
blinds 50—100 m from the nest tree at a location on a 
nearby cliff wall looking down on the nest with clear line- 
of-sight visibility. In one instance, it was not possible to 
build a blind with a direct view because of nest location, 
dense vegetation, or steep rock walls. In this instance, the 
nest was monitored from the ground at a similar dis- 
tance. Nest activities were recorded throughout the 
breeding season using spotting scopes, video cameras, 
and binoculars from the observation blind. The distance 
(m) between all occupied nests (nest-site spacing) for the 
2001 and 2002 breeding seasons was measured on the 
ground with surveying tape and verified on a study area 
map with Geographical Information System (GIS) mea- 
surements. A two-sample ^-test was used to determine if 
the spatial distribution of nest sites varied between years 
(Sheskin 2000). 

Continuous nest observations were conducted daily 
throughout the breeding season. Nest checks were ran- 
domly conducted throughout the day to include all hours 
when Broad-winged Hawks were active. Based on nest ob- 
servations and nest checks, we estimated dates of incu- 
bation, hatching, and fledging. We calculated nest sur- 
vival (Mayfield 1975) from start of incubation to fledging 
(total nest survival) and determined nest attentiveness 
patterns. A nest was considered successful if the pair pro- 
duced young. We estimated Mayfield nest survival during 
the incubation and nestling periods using a combined 
total of 198 incubation exposure-days and 274 nestling 
exposure-days. 

Prey Delivery. Food provisioning by adult Broad- 
winged Hawks to the nest during the breeding season was 
determined from direct observation. To assess if the birds 
regularly delivered prey throughout the day, occupied 
nests were monitored weekly in equal proportion during 
four time periods: early morning (0800-1100 H), early 
morning to mid afternoon (1101-1400 H), mid after- 
noon (1401—1700 H), and late afternoon to early evening 
(1701-2000 H). Nests were monitored from hatching un- 
til the young fledged or the nest failed. We calculated 
the mean number of prey items delivered and the pro- 
portion of prey deliveries per time period. We divided 
the prey items into two general categories, large (e.g., 
birds and rats) and small (e.g., macroinvertebrates and 
lizards) , to determine prey provisioning patterns. 


December 2005 


Puerto Rican Broad-winged Hawk Ecology 


407 


An analysis of variance in a randomized complete 
block design (PROC GLM; SAS Institute 1999) was used 
to test if Broad-winged Hawks delivered total number of 
prey items (response variable) per time period (block), 
number of large prey items per time period, and number 
of small prey items per time period equally throughout 
the day. We used multiple comparisons (Least Scientific 
Difference Means) of mean number of prey items 
(pooled, large, and small) to examine significant values 
(a = 0.05) and determine which time periods differed 
with respect to number of prey items delivered (Sheskin 
2000 ). 

Nest Habitat Model. Areas used by broadwings for 
nesting at Rio Abajo are valleys with tall forest bounded 
by limestone ridges and cliff walls, where pairs soar along 
their respective ridge tops (Delannoy and Tossas 2002). 
Forest vegetation along cliff walls and limestone ridges 
are used by resident broadwings for perching, but not 
for nesting (Hengstenberg and Vilella 2004). Our pri- 
mary objective was to assess which stand-level variables 
best described a broadwing nest within the context of the 
surrounding habitat at Rio Abajo Forest. Therefore, veg- 
etation characteristics and structure around nest sites 
were measured at the end of the breeding season and 
following post-fledging dependency. 

One nest used both in 2001 and 2002, was included 
once in the analyses. We described a nest site as all veg- 
etation within a 0.04 ha plot (11.3-m radius) centered on 
the nest tree (Titus and Mosher 1981). We recorded hab- 
itat measurements on nine of ten occupied nests and 
nine nonuse sites using standard procedures (James 
1971). We used a random numbers table to determine a 
distance and azimuth to travel from a particular nest site 
to an equivalent nonuse site. We constrained selection of 
random sites to forested area within a 400-m radius of 
the nest tree. All random sites were within the nest tree 
stand in valleys and side slopes. The closest tree to the 
plot center was chosen as the center point, and habitat 
variables were measured accordingly. 

We recorded visual obscurity of the understory using a 
2 m Nudds board (Nudds 1977). The board consisted of 
four 0.5 m sections with 30 orange and white squares. 
Nudds board measurements were taken from each car- 
dinal direction at a distance of 10 m from center point. 
Percentage visual obscurity for the four cardinal direc- 
tions was averaged for each 0.5 m section. We recorded 
altitude, aspect, percentage slope, and distance to the 
nearest rock wall, water, and man-made opening. All 
woody plants over 2 m tall according to species, diameter 
at breast height (DBH) , height, and vertical stratification 
were recorded. Vertical structure was classified into three 
strata heights (1-understory, 2-midstory, 3-overstory). 
Nest heights were recorded directly with a measuring 
tape. Tree heights were measured either by clinometer 
or through visual estimation. We tested for differences in 
height between clinometer readings and visual estima- 
tions using a two-sample West (SAS Institute 1999). We 
used a spherical densitometer to collect four readings of 
canopy cover from each cardinal direction at a distance 
of 5 m from center point. We calculated a mean to esti- 
mate percentage overstory canopy cover. 

Because of the paucity of information on nest site char- 
acteristics of Broad-winged Hawks in Puerto Rico (U.S. 


Fish and Wildlife Service 1997), we decided to record as 
much information as possible. We measured all variables 
considered biologically relevant to woodland raptors (Ti- 
tus and Mosher 1981). We tested habitat variables for 
normality using a Kolmogorov-Smirnov Goodness-of-Fit 
Test (Sheskin 2000). As data were normally distributed, 
we then used two-sample t-tests to identify variables that 
differed between nest sites and random sites. Uncorre- 
lated significant variables were selected for the variable 
selection model (see Table 1 for variables considered). 
Logistic regression analysis and AIC modeling was used 
to determine which variables best discriminate a Broad- 
winged Hawk nest site from a random site (PROC LO- 
GISTIC; SAS Institute 1999). 

To develop broadwing-habitat nest site relationships, 
we utilized a two model approach (variable selection and 
model selection) . For both microhabitat analyses, an in- 
formation-theoretic approach was used for model selec- 
tion and inference (Burnham and Anderson 2002). We 
used Akaike’s information criterion (AIC,.), delta AIC 
(A,), and Akaike’s ranking weight (w,) to determine the 
best model. 

Because of the small number of nest sites {N = 9), we 
recognized the variable selection model may produce bi- 
ased results (Milliken and Johnson 1984). Therefore, we 
conducted an alternative AIC model selection approach 
to assess which biologically-relevant variables best de- 
scribed Broad-winged Hawk nest sites. This alternative 
analysis models the particular nest sites without use of 
stand-level comparisons (nest site versus random site) as 
in the variable selection model. 

For the alternative model selection, we chose 10 ex- 
planatory variables from a list of 27 microhabitat vari- 
ables (Table 1) based upon literature review and person- 
al field experience (Titus and Mosher 1981, Delannoy 
and Tossas 2002). Variables chosen were: aspect, slope, 
road, rock wall, DBH, canopy cover, Nudds 2, midstory 
number of stems, overstory number of stems, and canopy 
height (Table 2). Nest or center tree height was excluded 
because it is significantly correlated with DBH. Best mod- 
el selection was based on criteria stated by Burnham and 
Anderson (2002). 

Results 

Breeding Behavior and Territorial Defense. 

From January to March of 2001 and 2002, we ob- 
served 158 courtship display flights by known pairs. 
Most (68%) aerial displays occurred 0917-1107 H 
{x = 1012 H). Across two breeding seasons, we doc- 
umented courtship display behavior in 26 pairs 
throughout the Rio Abajo Forest and surrounding 
private lands. Of the eleven pairs identified in 
2001, occupied nests were located for six of these. 
In 2002, courtship behavior was observed by 8 of 
the 11 pairs recorded in 2001. 

We observed 25 territorial interactions between 
Broad-winged Hawks and intruding Red-tailed 
Hawks. Broad-winged Hawks displaced Red-tailed 
Hawks 84% of the time when an intruding Red- 


408 


Hengstenberg and Vilella 


VoL. 39, No. 4 


Table 1. Nest habitat characteristics (mean ± SD, range) measured within 0.04 ha of Broad-winged Hawk nest and 
random sites in Rio Abajo Forest, Puerto Rico, 2001 and 2002. 


Habitat Characteristics 

Nest Site (9) 

Random Site (9) 

Mean ± SD 

Range 

Mean ± SD 

Range 

Altitude (m) 

235.7 ± 58.0 

(150-330) 

219.9 ± 62.7 

(126-320) 

Aspect 

204.9 ± 105.1 

(32-340) 

192.7 ± 108.8 

(12-334) 

Slope (%) 

46.0 ± 19.0 

(18-88) 

30.1 ± 24.3 

(0-85) 

Distance to water (m) 

194.1 ± 127.1 

(37-450) 

174.1 ± 120.7 

(3-370) 

Distance to road or trail (m) 

68.6 ± 35.5 

(25.8-135.0) 

40.8 ± 42.8 

(3.5-133.0) 

Distance to cliff wall (m) * 

41.1 ± 19.0 

(13-75) 

71.8 ± 30.2 

(35-137) 

Nest or center tree height (m)* 

22.2 ± 7.7 

(16.0-35.1) 

12.8 ± 5.2 

(6.5-22.0) 

Nest or center tree DBH (cm)* 

46.1 ± 15.6 

(23.0-74.5) 

25.2 ± 13.5 

(6.9-42.5) 

Nest height (m) 

16.3 ± 5.6 

(10.0-25.9) 


— 

Percentage nest height 

73.5 ± 6.7 

(58.8-81.3) 


— 

Canopy cover (%) 

85.2 ± 5.8 

(72-91) 

81.3 ± 6.0 

(70-88) 

Nudds 0.5 m (%) 

90.1 ± 11.5 

(63.7-100.0) 

75.9 ± 17.6 

(53.8-97.5) 

Nudds 1.0 m (%) 

75.8 ± 14.2 

(59.6-97.5) 

59.5 ± 25.4 

(21.6-87.6) 

Nudds 1.5 m (%)* 

74.4 ± 14.8 

(51.3-95.9) 

55.7 ± 18.0 

(25.8-77.7) 

Nudds 2.0 m (%)* 

77.1 ± 9.9 

(63.7-92.6) 

52.9 ± 24.5 

(15.9-83.5) 

Midstory species richness 

11.1 ± 5.6 

(3-18) 

12.0 ± 4.8 

(7-21) 

Midstory # of stems 

35.8 ± 16.7 

(10-58) 

58.6 ± 34.2 

(18-124) 

Midstory stem DBH 1-4.9 cm 

15.9 ± 12.7 

(1-36) 

27.4 ± 21.7 

(0-65) 

Midstory stem DBH 5-8.9 cm 

9.0 ± 6.1 

(2-19) 

16.0 ± 14.1 

(2-50) 

Midstory stem DBH s 9 

10.9 ± 5.3 

(7-22) 

16.3 ± 8.3 

(4-33) 

Overstory species richness 

3.1 ± 2.0 

(1-6) 

3.4 ± 1.9 

(2-7) 

Overstory # of stems 

10.2 ± 5.0 

(3-17) 

7.9 ± 5.1 

(2-16) 

Overstory stem DBH ^ 25.9 cm 

3.3 ± 2.9 

(0-9) 

3.0 ± 3.9 

(0-10) 

Overstory stem DBH 26-49.9 cm 

5.6 ± 3.2 

(2-11) 

4.2 ± 3.5 

(0-12) 

Overstory stem DBH ^ 50 cm 

1.3 ± 1.4 

(0-5) 

0.7 ± 0.9 

(0-2) 

Basal area m^/ha 

31.8 ± 13.6 

(14.2-53.4) 

28.2 ± 14.0 

(9.0-56.7) 

Canopy height (m) 

17.9 ± 2.8 

(14.4-23.4) 

16.6 ± 4.3 

(13-26) 


* Significant Kest (P < 0.05). 


tailed Hawk entered an occupied territory (p^ = 
21/25 = 0.84, P = 0.009, 2-tailed test). In every 
aerial encounter, the Red-tailed Hawk was the in- 
truding species. All aerial encounters involved 
adult birds of both species. Aerial displays involv- 
ing one Broad-winged Hawk and one Red-tailed 
Hawk occurred 72% of the time. However, Broad- 
winged Hawk pairs flew in unison 28% of the time 
to defend their territory against Red-tailed Hawks. 

Displays varied from “high-intensity flights” with 
intruders to “low-intensity flights” between pairs. 
During displays both birds circled together in close 
proximity and in the same general direction with 
the male broadwing flying above the female. Dur- 
ing low-intensity flights, adults soared upward on 
widespread wings and fanned tails. In high-inten- 
sity displays, the adults would alternate between 
wing flaps and soaring. When the male reached 
the top of the flight, he performed undulating 


dives or dipping flight (Wiley and Wiley 1981, 
Brown and Amadon 1989), consisting of a series of 
2-7 shallow dives made toward the female. Most 
other displays ended with the birds diving straight 
back with cupped wings at high speeds into the 
canopy, known as parachuting (Wiley and Wiley 
1981, Goodrich et al. 1996). 

Altitude gained varied among displays; generally 
the longer the flight, the higher the altitude ob- 
tained. Flights ranged from about 50 m above the 
tree canopy to elevations over 450 m. Courtship 
display flights lasted 60-900 sec (x = 334.7 ± 191.5 
sec). Cartwheeling behavior or sky dancing (Good- 
rich et al. 1996) was only observed three times by 
three different pairs, in which the male was flying 
on top of the inverted female with their talons ex- 
posed. The birds proceeded to tumble with semi- 
locked talons at high speeds until they reached the 
forested canopy, where they quickly released their 


December 2005 


Puerto Rican Broad-winged Hawk Ecology 


409 


Table 2. Model selection: parameters (K), relative AIC^, Delta AIC (A,), Akaike Weights (w,), Goodness-of-Fit 
statistic, P-value, and percent correct classification for Broad-winged Hawk nests in Rio Abajo Forest, Puerto Rico, 
2001 - 02 ). 


Model Selction 

K 

AIC, 

Ai 

w, 


P-VALUE 

Percent 

DBH + Canheight + Cliffwall + Overstems 

4 

20.057 

0.000 

0.304 

10.953 

0.027 

77.8 

DBH 

1 

21.362 

1.305 

0.158 

6.613 

0.010 

66.7 

Cliffwall 

1 

22.309 

2.252 

0.099 

5.283 

0.022 

72.2 

DBH + Canheight + Cliffwall 

3 

22.391 

2.334 

0.095 

9.009 

0.029 

72.2 

Nudds 2 

1 

22.399 

2.342 

0.094 

5.762 

0.016 

61.1 

DBH + Canheight 

2 

22.957 

2.900 

0.071 

7.167 

0.028 

66.7 

Aspect + DBH -1- Cliffwall 

3 

23.928 

3.871 

0.044 

8.569 

0.036 

72.2 

DBH + Canopy cover 

2 

24.619 

4.562 

0.031 

6.765 

0.034 

61.1 

Road + Cliffwall 

2 

25.591 

5.534 

0.019 

5.582 

0.061 

72.2 

Cliffwall + DBH 

2 

25.591 

5.534 

0.019 

8.493 

0.014 

72.2 

Midstems 

1 

26.050 

5.993 

0.015 

3.018 

0.082 

55.6 

Slope 

1 

26.974 

6.917 

0.010 

2.346 

0.126 

66.7 

Road 

1 

27.192 

7.135 

0.009 

2.211 

0.137 

55.6 

Canopy cover 

1 

27.435 

7.378 

0.008 

1.977 

0.160 

55.6 

Nudds 2 + Midstems + Overstems 

3 

27.961 

7.904 

0.006 

8.146 

0.043 

61.1 

Overstems 

1 

28.484 

8.427 

0.005 

1.023 

0.312 

59.0 

Midstems + Overstems 

2 

28.628 

8.571 

0.004 

3.486 

0.175 

50.0 

Canheight 

1 

28.862 

8.805 

0.004 

0.650 

0.420 

38.9 

DBH + Canheight + Cliffwall + Aspect 

4 

29.070 

9.013 

0.003 

9.047 

0.060 

66.7 

Aspect 

1 

29.458 

9.401 

0.003 

0.066 

0.797 

0.0 

Aspect + Slope 

2 

30.340 

10.283 

0.002 

2.395 

0.302 

72.2 


talon lock, swooped up, and dispersed upward in 
separate directions. 

Territorial flights were elicited by the presence 
or vocalizations from intruding Red-tailed Hawks 
and juvenile Broad-winged Hawks in the vicinity of 
the residents’ territory. Flights varied in intensity 
and depended on the intruding species and its 
proximity to the nest. Generally, males were first 
to fly and confront the intruder (Wiley and Wiley 
1981). Adults used alarm vocalizations to warn 
their mate of an intruding bird. During these dis- 
plays, adults used stuttered and whistle squeal vo- 
calizations (Burns 1911). Stuttered and whistle 
squeals vocalizations were used in high-intensity 
displays. When Red-tailed Hawks were detected. 
Broad-winged Hawks responded quickly with rapid- 
pursuit flights. Resident birds circled and soared 
to an altitude above the intruding bird and re- 
peatedly dived at it. The resident male continued 
to dive at the intruder until the intruder departed 
the territory. In some cases, the resident bird ex- 
tended its talons during its dives. In one instance, 
a male broadwing locked talons with an intruding 
Red-tailed Hawk; once the intruder left the area, 


the resident male silently dove or “parachuted” 
back to its territory. 

Dipping flight, or undulating display (Wiley and 
Wiley 1981, Brown and Amadon 1989), was a com- 
mon behavior used in all intense intruder inter- 
actions in which the resident bird was successful at 
chasing the Red-tailed Hawk. Territorial confron- 
tations between conspecific neighbors were less 
intense than toward Red-tailed Hawks. From the 
radiotelemetry study, a radio of a juvenile Broad- 
winged Hawk (5167) was found and all the feathers 
were plucked (D. Hengstenberg and F. Vilella un- 
publ. data). The cause of mortality was determined 
to be a Red-tailed Hawk, which had been observed 
numerous times in the same area as the juvenile 
broadwing. 

Perched intruders were generally attacked by a 
slower supplantation flight (Wiley and Wiley 1981) 
or a dive in which the intruder typically fled the 
area. If the intruder remained, the residents then 
circled above and used low angled dives until the 
intruder departed (Wiley and Wiley 1981). 

Nesting Biology. We found 10 nests during our 
study in Rfo Abtyo Forest. Onset of incubation was 


410 


Hengstenberg and Vilella 


VoL. 39, No. 4 


Table 3. Broad-winged Hawk nests monitored during 
the breeding seasons 2001 and 2002, Rio Abcgo Forest, 
Puerto Rico. 


Reproductive Variables 

2001 

2002 

Both 

Years 

Number of nests found 

6 

7 

13 

Occupied nests 

6 

4 

10 

Failed nests 

1 

2 

3 

Successful nests 

5 

2 

7 

Proportion successful nests 

0.83 

0.50 

0.70 

Mayfield nest success 

0.81 

0.51 

0.67 

Number of nestlings 

7 

3 

10 

Number fledged 

6 

2 

8 

Nestling loss 

0.14 

0.33 

0.20 

Fledglings per nest 

1.2 

1 

1.2 


from 28 February-21 March in 2001 and from 6- 
16 March in 2002. Hatching occurred from 9-20 
April in 2001 and from 6-17 April in 2002. In 2001, 
six juveniles fledged at 35-39 d between 2-25 May, 
and two juveniles fledged at 35-36 d from 21-22 
May in 2002. Three nests failed during the study 
(Table 3) . Two of the failed nests were attributed 
to heavy rains. The third nest was depredated by a 
Red-tailed Hawk. 

For 2001 and 2002, nest survival for the incu- 
bation and nestling periods was 0.67 (Table 3). 
The probability of surviving from nest initiation 
through fledging was 0.81 in 2001 and 0.51 in 
2002. For both years, the probability of nest surviv- 
al during the incubation period was 1.0. {N = 10). 
The number of fledglings per successful nest was 
1.20 in 2001, 1.0 in 2002, and 1.14 for both years 
combined (Table 1). 

During the egg stage, females spent the majority 
of time incubating. However, our results differ 
from the available information (Raffaele 1989, U.S. 
Fish and Wildlife Service 1997), as we documented 
males engaged in all nesting duties, including in- 
cubation. In some instances, males were observed 
continuously incubating for >4 hr and overnight. 
Adult Broad-winged Hawks frequently brought 
green vegetation, especially Trichilia hirta, to the 
nest. During incubation, vegetation may act as a 
buffer between nest branches and eggs. Raptors 
may use fresh greenery for concealment, to reduce 
odors, and to avoid ectoparasites (Wimberger 
1984, Sibley 2001). 

Nest Monitoring and Prey Delivery. We recorded 
5534 min of nest observations during the incuba- 


tion stage and 8825 min through the nestling 
stage. During the incubation period, females in- 
cubated approximately 53% of the time and males 
23%. We observed both male and female Broad- 
winged Hawks incubating overnight. The nest was 
not attended 15%, and an unknown adult incubat- 
ed 9% of the time. During the nestling period, the 
nest was not attended 69% of the time, females 
attended 17%, and males attended 14% of the 
time. As the nestlings matured, the female spent 
less time at the nest (Lyons and Mosher 1987). 

Prey deliveries away from the nest were relatively 
common, with either member of the pair initiating 
solicitation calls (Goodrich et al. 1996). The deliv- 
ering adult, usually the male, would vocalize back 
and forth until the incubating female flew off the 
nest to where the male was perched (<50 m from 
nest) to obtain the prey item. While one adult was 
eating, the other would fly to the nest and brood. 
Most prey deliveries during the nestling stage oc- 
curred during the early morning to mid-afternoon 
period. 

We observed 60 prey items delivered to 7 of 10 
monitored nests during the brood-rearing periods 
of 2001 and 2002 (Table 4). Prey consisted of 35% 
rats {Rattus , 27% lizards, 17% birds, 12% ma- 
croinvertebrates, 7% unidentified prey, and 3% 
snakes. Pooled prey items (large and small) varied 
among the four time periods (F^av = 4.01, P = 
0.024). More prey was delivered to nests during 
early morning to mid-afternoon than early even- 
ing. Daily prey deliveries were distributed as fol- 
lows: 38% early morning, 40% late morning to mid 
afternoon, 17% mid to late afternoon, and 5% late 
afternoon to early evening. However, neither the 
number of small prey items (F 327 = 2.83, P = 
0.068), nor the number of large prey items (^ 3,27 
= 2.80, P = 0.067) brought to nest sites, differed 
over the course of the day. The earliest prey deliv- 
ery was recorded at 0856 H, and the latest prey 
delivery was observed at 1846 H. We calculated a 
mean prey delivery rate of 0.38 prey (SE — 0.08) 
items per chick per hour (range = 0.14—0.80). 

Nesting Habitat. All nests were within 50 m of a 
rock wall. Nests sites were generally found on 
southwest facing slopes (x = 204°). Distance 
among nests averaged 838.5 m (SE = 79.98, range 
- 200-1455 m) in 2001 and 793.0 m (SE = 91.87, 
411-1231 m) in 2002. Distance did not vary be- 
tween years (^0 ~ 0.362, P = 0.720) . There was no 
difference amongst visual estimations and clinom- 
eter readings of nest tree heights when compared 


December 2005 


Puerto Rican Broad-winged Hawk Ecology 


411 


Table 4. Observed prey items delivered to Broad-winced Hawk nests or consumed in Rio Abajo Forest, Puerto Rico, 
2001 - 02 . 


Common Name 

Taxonomic Name 

Observed No. 
OF Prey 

Percent 

Puerto Rican giant centipede 

/ 

kjcuvujjt/vufu riuri '^ 

4 

6.7 

Puerto Rican arboreal millipede 

Orthocricus arboreus 

3 

5.0 

Melodius coqui 

Eleutherodactylus wightmanae 

1 

1.7 

Common coqui 

Eleutherodactylus coqui 

3 

5.0 

Common anole 

Anolis cristatellus 

2 

3.3 

Banded anole 

Anolis stratulus 

1 

1.7 

Yellow-breasted anole 

Anolis gundlachi 

2 

3.3 

Small green anole 

Anolis evermanni 

1 

1.7 

Orange dewlap anole 

Anolis krugi 

1 

1.7 

Snake anole 

Anolis pulchellus 

1 

1.7 

Green giant anole 

Anolis cuvieri 

2 

3.3 

Common gecko 

Spaherodactylus macrolepis 

2 

3.3 

Puerto Rican boa 

Epicrates inoratus 

1 

1.7 

Ground snake 

Arrhyton exiguum 

1 

1.7 

White-winged Dove 

Zenaida asiatica 

1 

1.7 

Bananaquit 

Coereba flaveola 

6 

10.0 

Puerto Rican Bullfinch 

Loxigilla portoricensis 

3 

5.0 

Common mouse 

Mus musculus 

12 

20.0 

Roof rat 

Rattus rattus 

7 

11.7 

Norway rat 

Rattus norvegicus 

2 

3.3 

Unidentified prey items 

— 

4 

6.7 

Total prey items 

— 

60 

100 


to actual tape-measured heights (P = 0.25). Nest 
tree height averaged 22.3 ± 7.7 m (range = 16.0- 

35.1 m) and nest height averaged 16.3 ± 5.6 m 
(range = 10.0-25.9 m). Nest tree DBH averaged 

46.1 ± 15.6 m (range = 23.0-74.5 cm). Dimen- 
sions for the two nests measured were: 0.79 m 
(long diameter) by 0.52 m (short diameter) by 0.61 
m (depth), and 0.46 m (long) by 0.31 m (short) 
by 0.61 m (depth). Nest cup depth measured 1.27 
and 2.54 cm, respectively. Within nest site vegeta- 
tion plots we recorded 13 species of overstory 
trees, but only four tree species were used as nest 
trees. Nests were in maria, Honduras mahogany, 
moca, and guaraguao trees. 

Of 27 microhabitat variables measured, five dif- 
fered between nests and random sites (Table 1). 
Nest sites were closer to cliff walls (t^Q = 2.578, P 
= 0.020), had greater tree height = —3.020, P 
= 0.008), larger DBH = -3.048, P = 0.008), 
and denser understories at 1.5 m (ti 5 = —2.409, P 
= 0.028) and 2.0 m = -2.742, P = 0.015) than 
random sites. 

Logistic regression produced a best nest site 
model containing two variables: DBH (parameter 


= 0.1800, SE = 0.0965, x^ = 3.4814, P = 0.062), 
and Nudds 2 m (parameter = 0.1297, SE = 0.0865, 
X^i = 2.2512, P = 0.134). This variable combina- 
tion correctly classified Broad-winged Hawk nests 
83.3% of the time (Table 5). The best AIC model 
for nest sites contained DBH (parameter = 1.2853, 
SE = 1.5429, xS = 0.694, P = 0.405), canopy 
height (parameter = 5.5472, SE = 7.5492, x^i ~ 
0.5399, P = 0.540), distance to rock wall (param- 
eter = -0.4281, SE = 0.8384, x^ = 0.2607, P = 
0.610), and overstory stems (parameter = 2.6233, 
SE = 3.5889, x^ = 0.5343, P = 0.465). These four 
variables correctly classified nests 77.8% of the 
time (Table 2). 

Discussion 

Albeit our small sample of nests, phenology was 
similar across both years, with the onset of incu- 
bation beginning in late February and juveniles 
fledging by the end of May. The post-fledging de- 
pendency period lasted 4—8 wk after the juveniles 
left the nest. During the first 2-3 wk post-fledging, 
juveniles frequently returned to the nest to receive 


412 


Hengstenberg and Vilella 


VoL. 39, No. 4 


Table 5. Variable selection of Broad-winged Hawk nest sites in Rio Abajo Forest, Puerto Rico, 2001-02. Significant 
parameters (K), relative AIC (AICJ, Delta AIC (AJ, Akaike Weight (w*), Goodness-of-Fit statistic (x^), P-value, and 
percent correct classification. 


Variable Selection 

K 

AIC, 

A, 



P-VALUE 

Percent 

DBH + Nudds 2 

2 

16.774 

0.000 

0.551 

10.754 

0.005 

83.3 

Cliffwall + DBH 

2 

19.386 

2.612 

0.149 

8.493 

0.014 

72.2 

Cliffwall + DBH -1- Nudds 2 

3 

19.630 

2.856 

0.132 

11.553 

0.009 

83.3 

DBH 

1 

21.362 

4.588 

0.056 

6.613 

0.010 

66.7 

Cliffwall + Nudds 2 

2 

21.770 

4.996 

0.045 

8.595 

0.014 

83.3 

Cliffwall 

1 

22.309 

5.535 

0.035 

5.283 

0.022 

72.2 

Nudds 2 

1 

22.399 

5.625 

0.033 

5.762 

0.016 

61.1 


prey deliveries from the adults and to roost for the 
night. 

Broad-winged Hawk nests in Puerto Rico aver- 
aged 1.1 young per nest attempt and 67% nest suc- 
cess. Our estimate is almost double what Delannoy 
and Tossas (2002) reported (0.66 fledglings/nest) 
for Broad-winged Hawk nests in Rio Abajo from 
1994 to 1996. However, our estimates of nest suc- 
cess and young per nest attempt were slightly lower 
compared to Broad-winged Hawk studies in North 
America (Armstrong and Euler 1983, Crocoll 1984, 
Rosenfield 1984). 

The number of fledglings per successful nest 
and overall nest success was greater in 2001 than 
in 2002. Lower nest success in 2002 may have been 
attributed to rain events that occurred in April of 
2002. Two nest sites were abandoned within the 
same week of heavy rain in April 2002. Santana and 
Temple (1988) reported lower success of Red- 
tailed Hawks nesting in the eastern Luquillo Moun- 
tains rainforest region during extensive rainy pe- 
riods. Similarly, severe rainfall was suggested as a 
cause of nest failures of the Puerto Rican Sharp- 
shinned Hawk {Accipiter striatus vennator) in the Lu- 
quillo Mountains (Snyder and Wiley 1976) and in 
forests of the central mountain range of the island 
(Delannoy and Cruz 1988). 

April and May are important months to nestling 
survival. During this time period, Broad-winged 
Hawks are brooding partially feathered nestlings. 
In 2001, total precipitation from April and May was 
44.2 cm. Conversely, April-May precipitation in 
2002 was 54.4 cm. April rains in 2002 coincided 
with the presence of recently-hatched chicks or 
young nestlings and may have caused nest aban- 
donment and hypothermia of young at the two 
failed nests. 

At Rio Abajo Forest, Broad-winged Hawks for- 


aged primarily on rats, lizards, and small birds (Ta- 
ble 4) . The Broad-winged Hawk is an opportunistic 
feeder who forages on a wide variety of prey (Rush 
and Doerr 1972, Reran 1978). However, our in- 
ability to detect a daily pattern of prey deliveries 
may have been a result of our small sample of nests 
and prey delivery observations. 

Prey size or type exploited at Rio Abajo Forest 
may be a function of seasonality (wet vs. dry), as 
changing weather conditions may produce differ- 
ences in dietary patterns (Grubb 1977, Stinson 
1980). In tropical environments such as Puerto 
Rico, rain may limit foraging opportunities (Foster 
1974). We observed Broad-winged Hawks at Rio 
Abajo Forest were less active during periods of 
rain. 

There was no difference in the spatial distribu- 
tion of nest sites between years, suggesting Broad- 
winged Hawks may maintain territories year round. 
These clusters of nests (Fig. 1) are bounded by 
limestone ridges and cliff walls, where pairs soar 
along their respective ridge tops. This may have 
some advantages. Pairs may be better able to detect 
intruding Red-tailed Hawks. Vigilance may contrib- 
ute to greater survival of nesting birds (Alcock 
1993). 

Breeding Broad-winged Hawks were aggressive 
and successfully deterred intruding Red-tailed 
Hawks from their nesting territories. Delannoy and 
Tossas (2002) speculated similar nest-site require- 
ments between Broad-winged Hawks and Red- 
tailed Hawks could lead to aggressive encounters. 
However, we found no evidence of nesting by Red- 
tailed Hawks within the closed canopy forests of 
Rio Abajo. By and large, Broad-winged Hawk court- 
ship and territory defense behavior in Puerto Rico 
was similar to that of the Ridgway’s Hawk {Buteo 


December 2005 


Puerto Rican Broau-winged Hawk Ecology 


413 


ndgwayi) in moist limestone forests of the Domin- 
ican Republic (Wiley and Wiley 1981). 

Territory occupancy seemed relatively stable be- 
tween years. The year-round residency and site fi- 
delity of Broad-winged Hawks in the moist lime- 
stone region of Puerto Rico may be indicative of 
long-term pair bonds (Griffin et al. 1998). Like 
other tropical raptors (Mader 1982, Griffin et al. 
1998), the subspecies in Puerto Rico seems to ex- 
hibit high site fidelity (73%, N — 11), 

Our results from radio-marked breeding birds 
(D. Hengstenberg and F. Vilella unpubl. data) sug- 
gest Broad-winged Hawks tend to nest in the same 
stand or nearby from one year to the next (Reran 
1978). We observed courtship behavior in 15 ad- 
ditional pairs, but found no evidence of nests or 
nest building. This may suggest that while some 
pairs may hold nesting territories, they do not nec- 
essarily build a nest or lay eggs every year, as has 
been documented in other raptors (Steenhof 
1987). 

Five nests constructed in maria trees were placed 
atop termite nests in the main crotch of the tree. 
Broad-winged Hawks nesting in North America 
sometimes place their nest on top of old bird and 
squirrel nests (Goodrich et al. 1996). 

Distance to cliff wall, tree height and DBH, 
Nudds board 1.5 m, and Nudds 2.0 m differed be- 
tween nest and random sites. Variables DBH and 
Nudds 2 m best classified nest sites, suggesting that 
Broad-winged Hawks at Rio Abajo Forest may pre- 
fer large trees and dense understories (Table 5). 
Also, Broad-winged Hawks in North America avoid- 
ed smaller trees and selected large DBH trees (Ti- 
tus and Mosher 1987). At Rio Abajo Forest, nest 
sites had denser understories than random sites. 
Dense understories may be related to prey avail- 
ability for adults. These dense understories may of- 
fer fledglings protection from predators and great- 
er foraging opportunities of prey. Radio-marked 
adults and juveniles were frequently observed 
hunting in the dense understory around their nest 
sites. Foraging habitat studies have suggested 
Broad-winged Hawks select sites with high prey 
availability (Steblein 1991). However, further re- 
search is required to better understand the rela- 
tionships between Broad-winged Hawks in Rio Aba- 
jo Forest and prey populations. 

Model selection procedures for nests (site spe- 
cific) yielded a four-variable model in which DBH, 
canopy height, cliff wall, and overstory stems cor- 
rectly classified nest sites (Table 2). This suggests 


closeness to karst cliff walls and canopy height may 
be additional predictors of Broad-winged Hawk 
nest habitat, in addition to basal area (i.e., DBH) 
and understory cover. 

Nest tree height averaged 22.2 m, whereas can- 
opy heights of nest plots averaged 17.9 m and ran- 
dom sites averaged 16.6 m, suggesting Broad- 
winged Hawks select emergent trees for their nests. 
Our results coincide with nest site characteristics 
of North American broadwings (Goodrich et al. 
1996). Delannoy and Tossas (2002) reported a 
mean nest tree height of 27.0 m and a mean can- 
opy height of 15.7 m. On average, nest heights in 
Puerto Rico were taller than nest heights reported 
from North America (Burns 1911, Matray 1974, Ti- 
tus and Mosher 1981, Armstrong and Euler 1983, 
Rosenfield 1984). This may reflect the relative 
lengths of trees in tropical versus temperate forests 
(Fedorov 1966). 

Broad-winged Hawk nest sites were located with- 
in 50 m of a cliff wall. In the karst region, cliff walls 
are very abundant. Cliff walls may offer nest sites 
with adequate protection from the elements (wind, 
rain, and sun), intruding predators, provide van- 
tage points, and facilitate reduction in energy re- 
quirements when searching for thermal updrafts. 
Nests sites were generally found on slopes facing 
southwest (x = 204°). This nest placement may 
help protect the nests from the prevailing easterly 
winds. Broad-winged Hawk nest sites in a limestone 
forest may be described as occurring in mature 
closed-canopy overstory stands sheltering a thin 
midstory, a dense understory, and in close prox- 
imity to a cliff wall. 

Conservation and persistence of the breeding 
population of Rio Abajo Forest may depend on 
management of the existing forest stands used by 
Broad-winged Hawks. Further research is required 
in Rio Abajo Forest to increase sample sizes of nest 
sites and validate the broadwing-habitat relation- 
ships revealed by our habitat model. Moreover, var- 
iance estimates of parameters from our study will 
provide baseline information needed to calculate 
sample sizes for future research. We suggest addi- 
tional studies to quantify broadwing habitat in oth- 
er localities of the karst region and to develop hab- 
itat models at multiple spatial scales. 

At Rio Abajo Forest, managers should limit dis- 
turbance within valleys used by broadwings during 
the critical nest initiation and incubation periods 
(i.e., February to April). Based on our preliminary 
results on habitat relationships, silvicultural prac- 


414 


Hengstenberg and Vtt.et.ia 


VoL. 39, No. 4 


tices within Rio Abajo Forest that promote main- 
tenance of canopy emergent trees and dense un- 
derstories may improve habitat conditions for 
nesting pairs as well as fledglings during their de- 
pendence period. Moreover, Broad-winged Hawks 
readily used plantation tree species such as maria 
and Honduras mahogany for nest sites. We rec- 
ommend the DNER Forest Service encourage sur- 
rounding private landowners to engage in agrofor- 
estry practices using these fast-growing plantation 
species. Additionally, programs for private lands 
that promote maintenance and enhancement of 
forest cover (e.g., USFWS Partners for Wildlife) 
should be brought to the attention of the land- 
owners adjoining Rio Ab^o Forest. 

In an attempt to establish a second wild popu- 
lation, releases of captive-reared Puerto Rican Par- 
rots are scheduled for 2006 (U.S. Fish and Wildlife 
Service 1999). Available information suggests 
Puerto Rican Parrots may exceed the size of avian 
prey taken by Broad-winged Hawks (Snyder and 
Kepler 1987). At Rio Abajo, 61% of prey deliveries 
to nests were rodents and Anolis lizards (Table 4) . 
Forest songbirds (e.g., Puerto Rican Bullfinch [Lo- 
xigilla portoricensis] and Bananaquit [ Coereba flaveo- 
la] ) were the avian prey taken (Table 4) . 

In contrast, Red-tailed Hawks are known parrot 
predators (White et al. 2005) . However, our results 
indicated resident Broad-winged Hawks chased off 
intruding Red-tailed Hawks effectively in Rio Abajo 
Forest. Owing to the likely negative relationship 
between these sympatric Buteos, resident Broad- 
winged Hawks in Rio Abajo may indirectly provide 
some degree of protection to released parrots from 
predation by excluding intruder Red-tailed Hawks. 
However, research is required to examine the re- 
lationship between spatial overlap of parrots and 
broadwings and the likelihood of Red-tailed Hawk 
predation on released parrots. 

Other studies have reported some avian species 
select nest sites close to more aggressive species 
that regularly attack or mob predators (Durango 
1949, Clark and Robertson 1979, Wiklund 1979, 
Dyrcz et al. 1981, Norrdahl et al. 1995). The Wood- 
pigeon {Columba palumbus) benefits from nesting 
in association with Eurasian (northern) Hobbies 
(Falco subbuteo; Bogliani et al. 1999). 

Nevertheless, both the Puerto Rican Parrot and 
Broad-winged Hawk are listed as endangered (U.S. 
Fish and Wildlife Service 1997, 1999). Therefore, 
a co-management approach will be required to en- 
sure habitat management activities for one species 


are not done at the expense of the other. We rec- 
ommend parrot habitat management activities 
(i.e., deployment of artificial cavities) should be 
limited to the nonbreeding season (August-De- 
cember) to minimize disturbance to Broad-winged 
Hawk nesting pairs and post-fledging dependent 
juveniles. 

Ultimately, the future of both these endangered 
species rests on the ability to disseminate research 
results to forest managers and policymakers. This 
information in turn will help to guide the protec- 
tion and conservation of the karst forest region of 
Puerto Rico, as further forest fragmentation will 
impact severely the recovery of both the broadwing 
and the parrot. Multiagency efforts are underway 
to acquire and protect a significant portion 
(>30 000 ha) of forest in the moist karst region of 
Puerto Rico (Lugo et al. 2001). Broad-winged 
Hawks do not limit their activities to the Rio Abajo 
Forest boundaries, and their fate in the surround- 
ing private lands may be uncertain. Therefore, 
DNER forest managers should work proactively 
with the surrounding land owners to promote 
land-use practices to conserve and to enhance ex- 
isting forest cover. Future patterns of land use 
around the forest boundary may indirectly and di- 
rectly affect the ability of the Rio Abajo Forest to 
function as an effective conservation unit for the 
Broad-winged Hawk. 

Acknowiedgments 

Funding was provided by the USFWS, Caribbean Field 
Office. The DNER Forestry Division provided access to 
Rio Abajo Forest. We are indebted to J. Casanova, DNER 
Management Officer at Rio Abajo Forest, and his staff 
for providing assistance and logistic support. The DNER 
Terrestrial Resources Division provided housing facilities 
at Rio Abajo Aviary. We are grateful to I. Llerandi, J. Rios, 
and A. Jordan for field assistance. Thanks to student vol- 
unteers from the University of Puerto Rico at Arecibo 
and Utuado. We are grateful for the helpful comments 
of C. Delannoy on the raptors of Puerto Rico. We thank 
J. Bednarz, K. Bildstein, L. Goodrich, K. Titus, and J. Wi- 
ley for suggestions that greatly improved earlier versions 
of the manuscript. Animal collection and handling pro- 
cedures were conducted under the auspices of DNER 
permit Ol-EPE-019, Master Station Permit 22456 of the 
USGS Bird Banding Laboratory, and Mississippi State 
University’s Institutional Animal Care and Use Commit- 
tee protocol No. 00-094. 

Literature Cited 

Ai,cock, j. 1993. Animal behavior: an evolutionary ap- 
proach. Sinauer Associates, Sunderland, MA U.S.A. 
Armstrong, E. and D. Euler. 1983. Habitat usage of two 


December 2005 


Puerto Rican Broad-winged Hawk Ecology 


415 


woodland Buteo species in central Ontario. Can. Field 
Nat. 97:200-207. 

Berger, D.D. and H.C. Mueller. 1959. The bal-chatri: a 
trap for the birds of prey. Bird Banding 30:18-2&. 

Bogliani, G., F. Sergio, and G. Tavecchia. 1999. Wood- 
pigeons nesting in association with Hobby Falcons: ad- 
vantages and choice rules. Anim. Behav. 57:125-131. 

Brown, L. and D. Amadon. 1989. Eagles, hawks, and fal- 
cons of the world. Wellflcct Press, Sec3.nciis, U.S.A. 

Burnham, K.P. and D.R Anderson. 2002. Model selection 
and multimodel inference. Springer-Verlag, New 
York, NY U.S.A. 

Burns, F.L. 1911. A monograph of the Broad-winged 
Hawk {Buteo platypterus) . Wilson Bull. 23:139-320. 

Cardona, J.E., M. Rivera, M. Vazquez Otero, and C.R. 
Laboy. 1987. Availability of food resources for the 
Puerto Rican Parrot and Puerto Rican Plain Pigeon 
in the Rio Abajo Forest. Final Report, Puerto Rico 
Department of Natural Resources, Project W-lO-ES-1. 

Clark, K.L. and R.J. Robertson. 1979. Spatial and tem- 
poral multi-species nesting aggregation in birds as 
anti-parasite and anti-predator defenses. Behav. Ecol. 
Sociobiol. 5:359-371. 

Clark, W.S. 1981. A modified dho-gaza trap for use at a 
raptor banding station. J. Wildl. Manage. 45:1043- 
1044. 

Crocoll, S.T. 1984. Breeding biology of Broad-winged 
and Red-shouldered Hawks in Western New York. 
M.S. thesis, State University College, Fredonia, NY 
U.S.A. 

and J.W. Parker. 1989. The breeding biology of 

Broad-winged and Red-shouldered Hawks in western 
New York. y. Raptor Res. 23:125-139. 

Daniel, W.W. 1990. Applied Nonparametric Statistics, 
2nd Ed. PWS-KENT Publishing Company, Boston, MA 
U.S.A. 

Delannoy, C.A. 1997. Status of the Broad-winged Hawk 
and Sharp-shinned Hawk in Puerto Rico. Carib.J. Sci. 
33:21-33. 

and a. Cruz. 1988. Breeding biology of the 

Puerto Rican Sharp-shinned Hawk {Accipiter striatus 
vennator). Auk 105:649—662. 

AND A.G. Tossas. 2002. Breeding biology and nest 

site characteristics of Puerto Rican Broad-winged 
Hawk at the Rio Abayo forest. Caribb. J. Sci. 38:20-26. 

Durango, S. 1949. The nesting association of birds with 
social insects and with birds of different species. Ibis 
91:140-143. 

Dyrcz, a., j. Witkowski, and J. Okulewicz. 1981. Nesting 
by ‘timid’ waders in the vicinity of ‘bold’ ones as an 
antipredator adaptation. Ibis 123:542—545. 

Erickson, M.G. and D.M. Hoppe. 1979. An octagonal bal- 
chatri trap for small raptors. Raptor Res. 13:36-38. 

Ewel, JJ- AND J.L. Whitmore. 1973. The ecological life 
zones of Puerto Rico and the U.S. Virgin Islands. 
USDA Forest Service Research Paper ITF-18. Rio 
Piedras, Puerto Rico. 


Federal Register. 1994. Endangered and threatened 
wildlife and plants; determination of endangered sta- 
tus for the Puerto Rican Broad-winged Hawk and the 
Puerto Rican Sharp-shinned Hawk. 59:46710—46715. 

Eedorov, A.A. 1966. The structure of the tropical rain 
forest and speciation in the humid tropics. J. Ecol. 54' 
1 - 11 . 

Pitch, H.S. 1974. Observations on the food and nesting 
of the Broad-y.'inged Hawk in northeastern Kansas. 
Condor 76:331—333. 

Poster, M.F. 1974. Rain, feeding behavior, and clutch 
size in tropical birds. Auk 91:722-726. 

Goodrich, L., J. Crocoll, and S.E. Senner. 1996. Broad- 
winged Hawk (Buteo platypterus ) . In A. Poole and F 
Gill [Eds.] . The birds of North America, No. 218. The 
Academy of Natural Sciences, Philadelphia, PAU.S.A., 
and the American Ornithologists’ Union, Washing- 
ton, DC U.S.A. 

Griffin, C.R., P.W.C. Paton, and T.S. Baskett. 1998. 
Breeding ecology and behavior of the Hawaiian 
Hawk. Condor 100:654-662. 

Grubb, T.C. 1977. Weather-dependent foraging in Os- 
preys. Auk 94:146-149. 

Hamerstrom, F. 1963. The use of Great Horned Owls m 
catching Marsh Hawks. Proc. XIII Int. Ornithol. Congr 
13:866-869. 

Hengstenberg, D.H. and FJ. ViLELLA, 2004. Reproduc- 
tive biology, abundance and movement patterns of 
the Broad-winged Hawk Buteo platypterus brunnescens in 
a moist limestone forest of Puerto Rico. Final Report, 
Cooperative Agreement No. 14-45-009-1543-59. U.S. 
Geological Survey Biological Resources Division, Mis- 
sissippi Cooperative Fish and Wildlife Research Unit. 
Mississippi State, MS U.S. A. 

James, F.C. 1971. Ordinations of habitat relationships 
among breeding birds. Wilson Bull. 83:215—236. 

Reran, D. 1978. Nest site selection by the Broad-winged 
Hawk in north-central Minnesota and Wisconsin. Rap- 
tor Res. 12:15-20. 

Lugo, A.E., L. Miranda Castro, A. Vale, T. del Mar 
Lopez, E. Hernandez Prieto, A. Garcia, A.R. Puente 
Rolon, A.G. Tossas, D.A. McEarlane, T. Millar, O 
Ramos, and E. Helmer. 2001. Puerto Rican karst — a 
vital resource. US Department of Agriculture Forest 
Service, International Institute of Tropical Forestry, 
GTR-WO-65. Rio Piedras, Puerto Rico. 

Lyons, D.M. and J.A. Mosher. 1987. Morphological 
growth, behavioral development, and parental care of 
Broad-winged Hawks. J. Field Ornithol. 58:334—344. 

Mader, W.J. 1982. Ecology and breeding habits of the 
Savanna Hawk in the llanos of Venezuela. Condor 84: 
261-271. 

Matray, P.F. 1974. Broad-winged Hawk nesting and ecol- 
ogy. Auk 91:307-324. 

Mayfield, H.F. 1975. Suggestion for calculating nest suc- 
cess. Wilson Bull. 87:456—466. 

Milliken, G.A. AND D.E. Johnson. 1984. Analysis of 


416 


Hengstenberg and Vilella 


VoL. 39, No. 4 


messy data. Vol. 1. Designed experiments. Van Nos- 
trand Reinhold Company, New York, NY U.S.A. 

National Oceanographic and Atmospheric Adminis- 
tration (NOAA) . 2002. Annual Climatological Sum- 
mary, National Climate Data Center. Asheville, 
NC U.S.A. www.ncdc.noad.gov/oa/climate/station/ 
stationlo cator. h tml 

Norrdahl, K., J. Suhonen, O. Hemminki, and E. Korpim- 
Aki. 1995. Predator presence may benefit: kestrels pro- 
tect curlew nests against nest predators. Oecologia 101: 
105-109. 

NcfDDS, T.D. 1977. Quantifying the vegetative structure of 
wildlife cover. Wildl. Soc. Bull. 5:113-117. 

Raffaele, H.A. 1989. A guide to the Birds of Puerto Rico 
and the Virgin Islands. Princeton University Press. 
Princeton, NJ U.S.A. 

Rosenfield, R.N. 1984. Nesting biology of Broad-winged 
Hawks in Wisconsin. Raptor Res. 18:6-9. 

Rush, D.H. and P.D. Doerr. 1972. Broad-winged Hawk 
nesting and food habitats. Auk 89:139-145. 

Santana, E.C., and S.A. Temple. 1988. Breeding biology 
and diet of Red-tailed Hawks in Puerto Rico. Biotropica 
20:151-160. 

SAS Institute. 1999. SAS/STAT® user’s guide, version 8, 
4th Ed., SAS Institute Inc., Cary, NC U.S.A. 

Sheskin, D.J. 2000. Handbook of parametric and non- 
parametric statistical procedures, 2nd Ed. Chapman 
and Hall, New York, NY U.S.A. 

Sibley, D.A. 2001. The Sibley guide to bird life & behav- 
ior. Alfred A. Knopf. New York, NY U.S.A. 

Snyder, N.F. and C.B. Kepler. 1987. The parrots of Lu- 
quillo: Natural history and conservation of the Puerto 
Rican Parrot. Western Foundation of Vertebrate Zo- 
ology, Los Angeles, CA U.S.A. 

AND J.W. Wiley. 1976. Sexual size dimorphism in 

hawks and owls in North America. Ornithol. Monogr. 

20 . 

Steblein, RF. 1991. Influence of spatial and temporal 
dynamics of prey populations on patch selection by 


Broad-winged Hawks (Abstract)./. Raptor Res. 25:160- 
161. 

Steenhof, K. 1987. Assessing raptor reproductive success 
and productivity. Pages 157-170 in B.G. Pendleton, 
B.A. Millsap, K.W. Cline, and D.M. Bird [Eds.], Insti- 
tute for Wildlife Research. Raptor Management Tech- 
niques Manual. Scientific and Technical series No. 10, 
National Wildlife Federation, Washington, DC U.S.A. 

Stinson, C.H. 1980. Weather-dependent foraging success 
and sibling aggression in Red-tailed Hawks in central 
Washington. Condor 82:76—80. 

Titus, K. 1987. Selection of nest tree species by Red- 
shouldered and Broad-winged hawks in two temper- 
ate forest regions. J. Field Ornithol. 58:274—283. 

AND J.A. Mosher. 1981. Nest-site habitat selected 

by woodland hawks in the central Appalachians. Auk 
98:270-281. 

U.S. Fish and Wildlife Service. 1997. Puerto Rican 
Broad-winged Hawk and Puerto Rican Sharp-shinned 
Hawk Recovery Plan. U.S. Fish and Wildlife Service, 
Atlanta, GA U.S. A. 

. 1999. Technical/Agency Draft Revised Recovery 

Plan for the Puerto Rican Parrot {Amazona vittata). 
U.S. Fish Wildlife Service, Atlanta, GA U.S.A. 

Vekasy, M.S., J.M. Marzluff, M.N, Kochert, R.N. Leh- 
man, and K. Steenhof. 1996. Influence of radio trans- 
mitters on Prairie Falcons, f. Field Ornithol. 67:680- 
690. 

White, T.H., J.A. Collazo, and F.J. Vilella. 2005. Sur- 
vival of captive-reared Puerto Rican Parrots released 
in the Caribbean National Forest. Condor 107:424— 
432. 

Wiklund, C.G. 1979. Increased breeding success for Mer- 
lin {Falco columbarius) nesting among fieldfare (Turdus 
pilaris) colonies. Ibis 121:109-111. 

Wiley, J. A. and B.N. Wiley. 1981. Breeding season ecol- 
ogy and behavior of Ridgway’s Hawk (Buteo ridnuayi). 
CowtZor 83:132-151. 

Wimberger, P.H. 1984. The use of green plant material 
in bird nests to avoid ectoparasites. Auk 101:615-618. 

Received 28 June 2004; accepted 27 August 2005 


J. Raptor Res. 39(4):417-428 
© 2005 The Raptor Research Foundation, Inc. 

RAPTOR ABUNDANCE AND DISTRIBUTION IN THE LLANOS 

WETLANDS OF VENEZUELA 

Wendy J. Jensen, 1 Mark S. Gregory, and Guy A. Baldassarre 

State University of New York, College of Environmental Science and Forestry, 1 Forestry Drive, 

Syracuse, NY 13210 US. A. 

Francisco J. Vilella 

uses Biological Resources Division, Cooperative Research Units, Department of Wildlife and Fisheries, 

Box 9691, Mississippi State, MS 39762-9691 U.S.A. 

Keith L. Bildstein 

Acopian Center for Conservation Fearning, Hawk Mountain Sanctuary, 410 Summer Valley Road, 

Orwigsburg, PA 17961 U.S.A. 

Abstract. — The Llanos of Venezuela is a 275 OOO-km® freshwater wetland long recognized as an impor- 
tant habitat for waterbirds. However, litde information exists on the raptor community of the region. 
We conducted raptor surveys in the Southwestern and Western Llanos during 2000-02 and detected 28 
species representing 19 genera. Overall, areas of the Llanos that we sampled contained 52% of all raptor 
species and more than 70% of the kites, buteos, and subbuteos known to inhabit Venezuela. Regional 
differences in the mean number per route for four of the 14 most common species, the Crested Ca- 
racara ( Caracara plancus ) , Black-collared Hawk {Busarellus nigricollis) , American Kestrel {Falco sparverius) , 
and Osprey (Pandion haliaetus), were significant {P < 0.0018) in relation to the wet or dry seasons. Of 
the 14 less common species, six were detected in only one season (wet or dry). The Southwestern and 
Western regions of the Llanos support a rich raptor community composed primarily of nonmigratory 
wetland-dependent and upland-terrestrial species. 

Key Words: Neotropics-, Venezuela, Llanos, savanna, wetlands', roadside surveys. 


DISTRIBUCION YABUNDANCIA DE RAPACES EN HUMEDALES DE LOS LLANOS DE VENZUELA 

Resumen. — Los llanos de Venezuela constituyen un humedal de agua dulce de 275 000 km^ que ha sido 
tradicionalmente reconocido como un ambiente importante para las aves acuaticas. Sin embargo, existe 
poca informacion sobre la comunidad de rapaces de la region. Realizamos censos de aves rapaces en 
el sudoeste y el oeste de los llanos entre 2000 y 2002 y detectamos 28 especies que representaron 19 
generos. En total, las areas de los llanos que censamos contuvieron el 52% de todas las especies de 
rapaces y mas del 70% de los elanios, buteos y subbuteos que habitan en Venezuela. Las diferencias 
regionales en el numero medio por ruta para cuatro de las 14 especies mas comunes, Caracara plancus, 
Busarellus nigricollis, Falco sparverius y Pandion haliaetus, fueron significativas (P< 0.0018) con relacion a 
las estaciones humeda y seca. De las 14 especies menos comunes, seis fueron detectadas en una sola 
estacion (humeda o seca). Las regiones del sudoeste y del oeste de los llanos albergan una rica comu- 
nidad de aves rapaces compuesta primariamente por aves no migratorias que dependen de humedales 
y de especies terrestres de lugares elevados. 

[Traduccion del equipo editorial] 


South America comprises 12% of the world’s 
land surface, yet supports 28% of all raptors (Bie- 
rregaard 1998). Most South American raptors do 
not appear threatened globally, but more infor- 
mation is needed to confirm current assessments 


^ Email address: wjjensen@syr.edu 


and appropriately address threats (Bierregaard 
1998, Bildstein et al. 1998). Community level rap- 
tor research in South America has been primarily 
focused in forest habitats (e.g., Thiollay 1984, 
Thiollay 1989, Alverez et al. 1996, Manosa and Pe- 
drocchi 1997, Manosa et al. 2003); thus, little is 
known about the raptor populations within the ex- 


417 


418 


Jensen et al. 


VoL. 39, No. 4 


tensive savanna and grassland regions of the con- 
tinent. Raptor surveys in nine South American 
countries during 1979 detected the greatest num- 
ber of species in the savannas, mixed riparian for- 
ests, pastures, and open areas of interior Venezuela 
(Ellis et al. 1990), indicating that these areas are 
important habitats for South American raptors. 
Furthermore, Neotropical raptors of open land 
and savanna habitats are currently threatened by 
habitat loss, including wetland depletion and land- 
scape homogenization (Alvarez-Lopez and Kattan 
1995). 

In Venezuela, the savannas of the interior are 
part of an extensive (275 000 km^) wetland com- 
plex called the Llanos. The Llanos cover approxi- 
mately 31% of Venezuela (Mittermeier et al. 2003) 
and are located in the latitudinal region character- 
ized by the greatest avian endemism in the North- 
ern Hemisphere (Bibby et al. 1992). Between 32- 
36 nonmigratory and North American migratory 
raptor species use some or all of the Llanos (Fer- 
guson-Lees and Christie 2001, Hilty 2003). Al- 
though the natural history, biology, and habitat as- 
sociations of some of these species have been 
studied locally (Mader 1981, 1982, Beissinger et al. 
1988, Balgooyen 1989, Kirk and Currall 1994), 
community based, landscape-level surveys are lack- 
ing. 

Our objective was to document and compare the 
species richness, relative abundance, and distribu- 
tion of nonmigratory and migratory raptors in the 
savannas of the Southwestern and Western regions 
of the Venezuelan Llanos. We also compared these 
population parameters between the distinct wet 
and dry seasons that characterize the Llanos. 

Study Area 

Venezuela supports 1381 species of birds (Hilty 2003) 
and is considered a globally important region of biodi- 
versity in part due to its rich avifauna (Mittermeier and 
Mittermeier 1997, Myers et al. 2000), The Venezuelan 
Llanos is located between ca. 6-9°N and 63— 71 TV and is 
bordered by the Coastal Cordillera to the north, the Ori- 
noco Delta and Guiana Shield to the east and southeast, 
Colombia to the south and southwest, and by the Andes 
Mountains to the northwest. 

Annual rainfall in the Venezuelan Llanos ranges from 
90-180 cm (Silva and Moreno 1993), with most rain fall- 
ing and widespread flooding occurring from April 
through November (Cole 1986). In contrast, the late No- 
vember-late April dry season typically is rain free (Troth 
1979). 

The Venezuelan Llanos is divided into three general 
areas: western, central, and eastern (Huber and Alcaron 
1988). Covering 90 000 km‘^, the western area comprises 


35% of the freshwater wetland in Venezuela and spans 
north and west of the Orinoco River from ca. 69-7l°W 
and 6-9°N (Bulla et al. 1990). This area is relatively flat, 
with elevation ranging from sea level at the Orinoco Riv- 
er to 155 m near the foothills of the Andes. The western 
area is further divided into two distinct regions known as 
the Southwestern and Western Llanos (Fig. 1; Huber and 
Alcaron 1988). 

The Southwestern Llanos includes the vast open-sa- 
vanna-wetland habitats that extend from the Meta and 
Orinoco rivers northwest to the agricultural-savanna-for- 
est mosaic habitats of the Western Llanos. The South- 
western Llanos is characterized by poor soils, savanna 
habitats with small patches of trees, gallery forest, and 
extensive wet season flooding that renders the region 
largely unsuitable for agriculture (Huber and Alcaron 
1988). 

The Western Llanos encompasses the alluvial plains 
bordered by the foothills of the Andes and extends south- 
east to the open savannas of the Southwestern Llanos. 
This region is characterized by fertile soils and partial 
flooding that support agricultural production and native 
forests dominated by tree species similar to those of the 
Amazon basin (Huber and Alcaron 1988, Silva and Mo- 
reno 1993). 

Methods 

Sampling Design. We surveyed raptors along the sparse 
network of roads accessible during both the wet and dry 
seasons in the Southwestern and Western Llanos (Fig. 1) 
Despite limitations inherent in roadside surveys (Millsap 
and LeFranc 1988, Bunn et al. 1995), such counts can be 
used to survey relative raptor abundance, community 
composition, and habitat associations across large land- 
scapes (Woffinden and Murphy 1977, Thiollay 1978, Ellis 
et al. 1990, Sorley and Andersen 1994, Seavy and Apo- 
daca 2002). However, roadside surveys in open habitats 
may be inadequate for detecting small and uncommon 
raptor species unless surveys incorporate frequent and 
regular stops (Whitacre and Turley 1990). Therefore, we 
used the North American Breeding Bird Survey model 
(Droege 1990) to establish stationary surveying points 
along road routes. 

We placed 50 survey routes along roads where the over- 
all distribution of survey routes was dictated by accessi- 
bility of roads during the wet season (Fig. 1). Each route 
was 22.5 km long with 16 sample points spaced 1.5 km 
apart (Jensen 2003: Appendix A) . Each sample point was 
surveyed by a 2-person surveying team for 3 min to the 
right and left side of the road for a total of 6 min/ point. 
All raptors were tallied within a 500-m radius as measured 
by rangefinder binoculars. Except for the King Vulture 
{Sarcoramphus papa), we did not tally Cathartid vultures. 
Scientific and common names of birds follow Ferguson- 
Lees and Christie (2001). 

Raptors were surveyed during rainless periods of the 
day, primarily from 0700-1200 H, and never later than 
1400 H. Surveys were conducted over ca. 6-wk periods 
twice each year between August 2000 and March 2002 
The 6-wk periods coincided with the end of the wet (Au- 
gust-October) and dry seasons (January-March). We 
surveyed 27 routes in the wet season of 2000, 39 in the 


December 2005 


The Raptor Community of the Llanos 


419 




North Central Llanos 


Central 


Western Llanos 


Llanos 


Southwestern Llanos 


Legend 

c j Regions of the Venezuelan Llanos 
n The Colonijian L i^nos 


crwTis 


10°N 


8"N 


6"N 


70" W 


68" W 


66"Vy 


Figure 1. Study area map of the raptor survey during wet seasons 2000 and 2001 and dry seasons 2001 and 2002 
in the Southwestern and Western Llanos of Venezuela. Route clusters are areas where groups of survey routes con- 
taining sample points are located. 



420 


Jensen et al. 


VoL. 39, No. 4 


dry season of 2001, 50 in the wet season of 2001, and 50 
in the dry season of 2002 . 

Analysis. We calculated the total number of individuals 
detected on all surveys, as well as species totals, percent 
composition, and frequency of detection. We classified 
percent composition into four species abundance classes: 
Very common (10-26% of all individuals detected), Com- 
mon (3-5%), Uncommon (1-2%), and Rare (<1%). We 
used Estimates Version 6 . Obi software (Colwell 2000) to 
generate a species accumulation curve to evaluate the 
probability that our 2-yr 50-route roadside survey design 
was adequate for documenting all detectable raptor spe- 
cies. 

We then calculated the mean number of individuals 
per species per route (mean number per route) by year, 
season, and region to investigate annual changes in abun- 
dance between the 2000 and 2001 wet seasons and the 
2001 and 2002 dry seasons for each region. For this com- 
parison we used the 27 routes surveyed during the first 
survey visit (wet season 2000 ) and resurveyed for the du- 
ration of the survey (2001 and 2002): Southwestern Lla- 
nos (15 routes) and Western Llanos (12 routes). Rare 
abundance class species were detected in numbers too 
low to meet the underlying assumptions for a /-test and 
were omitted from all subsequent statistical tests. For the 
14 most common species, we used paired /-tests to eval- 
uate four hypothesis (Hoj: mean number per route 
Southwestern Llanos wet season 2000 = mean number 
per route Southwestern Llanos wet season 2001; Hog: 
mean number per route Southwestern Llanos dry season 
2001 = mean number per route Southwestern Llanos dry 
season 2002; Hog: mean number per route Western Lla- 
nos wet season 2000 = mean number per route Western 
Llanos wet season 2001; H 04 : mean number per route 
Western Llanos dry season 2001 = mean number per 
route Western Llanos dry season 2002). Differences were 
considered significant at L* < 0.10 because we were in- 
terested in large-scale broad patterns. However, we used 
the Bonferroni Method to control for inflated experi- 
mentwise type I error rate resulting from simultaneous 
multiple comparisons (Beal and Khamis 1991). For all 
analysis, Bonferroni corrected significance for 56 com- 
parisons was determined at {P < 0.0018). 

We then combined data from both years to investigate 
seasonal and regional differences in species numbers, 
species diversity, community composition, and mean 
number per route for each species. During the 2-yr study, 
50 routes were surveyed (Southwestern Llanos 28 routes, 
Western Llanos 22 routes) . We surveyed 27 routes during 
both survey years (wet season 2000 and dry season 2001 ) 
and an additional 23 routes during the second survey 
year (wet season 2001 and dry season 2002). To stan- 
dardize the number of individuals detected on the 27 
routes surveyed over 2 -yr, we averaged the number of 
individuals per survey point across survey years for each 
species. We then combined this 27-route average with the 
data from the 23 routes surveyed only during the second 
survey year to yield 50 total routes. Comhining the data 
using this method preserved the majority of data collect- 
ed, accounted for variation in numbers of individuals de- 
tected for each species on routes replicated over survey 
years, and preserved data from routes surveyed only dur- 
ing the second year. 


We calculated species numbers and Simpson’s Inverse 
Diversity Index (D = 1/X pi^) for the wet and dry seasons 
within each region (Hayek and Buzas 1996). We also cal- 
culated Jaccard’s Coefficient of Community Similarity to 
estimate the percent overlap between communities in 
both seasons and regions (Magurran 1988). Again, we 
used Estimates version 6.0b 1 software (Colwell 2000) to 
generate species accumulation curves for each regional 
and seasonal dataset to evaluate the probability that the 
number of routes surveyed in each region for each sea- 
son were adequate for documenting all detectable raptor 
species. 

To investigate seasonal and regional differences in 
numbers for each species, we first calculated the mean 
numbers per route for each region and season. To com- 
pare seasonal di ff erences in mean numbers per route for 
the 14 most abundant species, we used paired /-tests to 
evaluate two hypothesis (Hop mean number per route 
Southwestern Llanos wet seasons = mean number per 
route Southwestern Llanos dry seasons; Hog: mean num- 
ber per route Western Llanos wet seasons = mean num- 
ber per route Western Llanos dry seasons). For the same 
14 species, we also used 2-sample /-tests to evaluate re- 
gional differences in mean numbers per route (Hop 
mean number per route Southwestern Llanos wet sea- 
sons = mean number per route Western Llanos wet sea- 
sons; Hog: mean number per route Southwestern Llanos 
dry seasons = mean number per route Western Llanos 
dry seasons). The 14 rare abundance class species are 
presented only as present or absent for each region and 
season. 

Results 

General Patterns. We counted 5735 raptors rep- 
resenting 28 species and 19 genera (Table 1). The 
four most abundant species, Crested Caracara ( Ca- 
racara plancus). Yellow-headed Caracara (Milvago 
chimachima) , Savanna Hawk {Buteogallus meridiona- 
lis), and Roadside Hawk (Buteo magnirostris) , com- 
prised 72% of all individuals and were seen on 98- 
100% of all routes. Four additional species were 
classified as common and together comprised 15% 
of all individuals. These were the Black-collared 
Hawk {Busarellus nigricollis) , White-tailed Hawk 
{Buteo albicaudatus) , Snail Kite {Rostrhamus sociabi- 
lis) , and American Kestrel {Falco sparverius) . Six ad- 
ditional species were uncommon and represented 
10% of all detections. They were the Aplomado 
Falcon {Falco femoralis) , Great Black Hawk {Buteo- 
gallus urubitinga), White-tailed Kite {Elanus leucu- 
rus) , Crane Hawk ( Geranospiza caerulescens) , Laugh- 
ing Falcon {Herpetotheres cachinnans), and Osprey 
{Pandion haliaetus). Fourteen additional species 
comprised the remaining 3%. Species classified as 
common were detected on 52-84% of all routes, 
whereas uncommon species were detected on 44- 
78% routes. The 14 rare species were detected on 


December 2005 


The Raptor Community of the Lianos 


421 


Table 1. Species detected during raptor surveys in the Southwestern and Western Llanos of Venezuela during wet 
seasons in 2000 and 2001 and dry seasons in 2001 and 2002. Species are listed in order of relative abundance based 
on percent composition. English common names and taxonomy follow Ferguson-Lees and Christie (2001). 


Relative 

Abundance 

Class"' 

Species 

Status'^ 

Total 

Number 

Percent 

Composition 

Frequency 

Occurrence 

(%)<= 

Very common 

Creasted Caracara ( Caracara plancus) 

R 

1473 

26 

100 


Yellow-headed Caracara {Milvago chimachima) 

R 

1061 

19 

100 


Savanna Hawk {Buteogallus meridionalis) 

R 

1005 

18 

98 


Roadside Hawk {Buteo magnirostris) 

R 

604 

11 

98 

Common 

Black-collared Hawk {Busarellus nigticollis) 

R 

269 

5 

84 


Snail Kite {Rostrhamus sodabilis) 

R 

226 

4 

68 


American Kestrel {Falco sparverius) 

R/NAM 

183 

3 

52 


White-tailed Hawk {Buteo albicaudatus) 

R 

155 

3 

72 


Aplomado Falcon {Falco femor alls) 

R 

113 

2 

78 

Uncommon 

Great Black Hawk {Buteogallus urubitinga) 

R 

110 

2 

58 


White-tailed Kite {Elanus leucurus) 

R 

137 

2 

70 


Crane Hawk {Geranospiza caerulescens) 

R 

74 

1 

56 


Laughing Falcon {Herpetotheres cachinnans) 

R 

70 

1 

44 


Osprey {Pandion haliaetus) 

NAM 

66 

1 

48 

Rare 

King Vulture ( Sarcoramphus papa) 

R 

41 

0.7 

24 


Harris’s Hawk {Parabuteo unidnctus) 

R 

36 

0.5 

28 


Grey-lined Hawk {Buteo nitidus) 

R 

27 

0.5 

28 


Zone-tailed Hawk {Buteo albonotatus) 

R 

19 

0.3 

32 


Slender-billed Kite {Rostrhamus hamatus) 

R 

19 

0.3 

16 


Bat Falcon {Falco rufigularis) 

R 

16 

0.3 

24 


Plumbeous Kite {Ictinia plumbea) 

R 

9 

0.2 

14 


Hook-billed Kite ( Chondrohierax undnatus) 

R 

5 

0.08 

10 


Long-winged Harrier {Circus buffoni) 

R 

5 

0.09 

6 


Common Black Hawk {Buteogallus anthradnus) 

R 

5 

0.09 

10 


Peregrine Falcon {Falco peregrinus) 

NAM 

5 

0.09 

8 


Short-tailed Hawk {Buteo brachyurus) 

R 

5 

0.09 

6 


Grey-headed Kite {Leptodon cayanensis) 

R 

3 

0.05 

4 


Pearl Kite ( Gampsonyx swainsonii) 

R 

3 

0.05 

6 


® Relative abundance class: Very common = 10-26% of all individuals detected; Common = 3-5%, Uncommon = 1-2%, Rare = 
< 1 %. 

^ Status: R = nonmigratory population, NAM = North American migratory population, R/NAM = nonmigratory and North American 
migratory populations. 

Frequency of occurrence: the percent of routes on which a species was detected. 


<32% of all routes. There were three species of 
North American migrants: Osprey, American Kes- 
trel, and Peregrine Falcon {Falco peregrinus). The 
Osprey and Peregrine Falcon combined comprised 
1.2% of all individuals seen. In contrast, the Amer- 
ican Kestrel comprised 3%, having both nonmigra- 
tory and migratory populations (Hilty 2003). At 
least one of these three species was seen at least 
once on 82% of routes. 

The species accumulation curve indicated the 
number of routes surveyed over the two survey 
years was adequate for documenting all species de- 
tectable by roadside point count surveys in the 


study area (Fig. 2) . Specifically, all 28 species were 
detected with 34 routes (68% of all routes sur- 
veyed) . 

Yearly Comparisons. Among the 14 most com- 
mon species there were no regional or seasonal 
differences (P > 0.0018) in mean number per 
route between survey years (Table 2). 

Regional and Seasonal Patterns. The greatest 
number of species, 25, was detected in the South- 
western Llanos during the wet season and in the 
Western Llanos during the dry season (Fig. 3) . Al- 
though species numbers were equal between re- 
gions, diversity was higher in the Western Llanos 


422 


Jensen et al. 


VoL. 39, No. 4 



Nutnbar of Routes Surveyed 


Figure 2. Species accumulation graph of raptor species 
detected on 50 survey routes in the Southwestern and 
Western Llanos of Venezuela during the wet seasons in 
2000 and 2001 and dry seasons 2001 and 2002. 

during both seasons. Species accumulation curves 
indicated the number of routes surveyed during 
each season was adequate to detect the majority of 
species for both regions (Fig. 4) . 

The raptor community (Jaccard’s Coefficient of 
Community Similarity) in the Southwestern versus 
Western Llanos differed by 12%, both in the wet 
and dry seasons. For seasons combined, 25 species 
were detected in each region, of which 22 species 
were shared between regions. The seasonal chang- 
es in community composition within regions 
(26%) were greater than the regional differences 
within the wet and dry season (12%). 

Four of the 14 most common species exhibited 
regional differences {P< 0.0018) in mean number 
per route in relation to the wet or dry seasons (Ta- 
ble 3). The Crested Caracara (P < 0.001) and Os- 
prey (P < 0.001) were more numerous in the 
Southwestern Llanos than the Western Llanos dur- 
ing the wet season. The Black-collared Hawk (P = 
0.001) was more numerous in the dry season in 
the Southwestern Llanos than the Western Llanos, 
whereas the American Kestrel (P = 0.001) was 
more numerous in the dry season in the Western 
Llanos than the Southwestern Llanos. 

Of the 14 rare species, six were detected only 
during one season or region (Table 4). The Pere- 
grine Falcon and Grey-headed Kite were seen only 
in the dry season, and the Pearl Kite and Slender- 
billed Kite were seen only in the wet season. The 
Common Black Hawk was detected only in the 
Southwestern Llanos in the wet season, and the 
Plumbeous Kite was detected only in the Western 
Llanos in the dry season. 


Discussion 

General Findings. The savannas of the South- 
western and Western Llanos of Venezuela are par- 
ticularly rich in raptors, supporting 52% (32 of 61) 
of all regularly occurring migrant and resident spe- 
cies found in Venezuela (Hilty 2003) . Indeed, dur- 
ing our 2-yr study of the Venezuelan Llanos, these 
regions included 55% of all hawk species (10 of 
18), 70% (7 of 10) of kite species, 67% (4 of 6) of 
vulture species, and 50% (3 of 6) of the regularly 
occurring North American migratory species 
(Hilty 2003). However, although the Llanos sup- 
ported species assemblages equivalent to all other 
Venezuelan life zones for most raptors, we did not 
detect any of the eight eagle species that occur in 
Venezuela. 

Except for the three common Cathartid species 
we did not count, we detected 24 of 27 resident 
species expected to occur in the Llanos (Hilty 
2003) . We did not detect three uncommon forest- 
dwelling raptors thought to occur in the region: 
the Collared Forest-Falcon (Micrastur semitorqua- 
tus). Bicolored Hawk {Accipiter bicolor), and Ornate 
Hawk Eagle (Spizaetus ornatus). We likely failed to 
see these species because of the inherent difficulty 
of detecting forest-dwelling species from roadside 
surveys (Millsap and LeEranc 1988). 

Of the three North American migrant species we 
detected, two occur year-round in the Llanos. The 
American Kestrel occurs year-round in the Llanos 
because there are nonmigratory and migratory 
populations (Hilty 2003) . Although the Osprey 
population is wholly migratory, the Osprey occurs 
year-round in the Llanos because first-year birds 
reaching the Llanos remain for at least 18 mo 
(Martell et al. 2001, Hilty 2003). 

North American migratory species not seen dur- 
ing surveys were the Northern Harrier {Circus cya- 
neus), Broad-winged Hawk {Buteo platypterus) , 
Swainson’s Hawk {B. swainsoni), and Merlin {Falco 
columbarius) . However, one Merlin was detected in 
the Llanos, but not on survey routes. Historical 
sightings of the Swainson’s Hawk and Northern 
Harrier in the Llanos are considered accidental 
(Hilty 2003), and satellite-tracking of Swainson’s 
Hawks confirms that their migrations to and from 
Argentina occur along the central and eastern 
slopes of the Andes (Euller et al. 1998) . The Broad- 
winged Hawk is thought to winter in portions of 
the Llanos west and north of our study area (Hilty 
2003). 


December 2005 


The Raptor Community of the Llanos 


J= 


U 

> 

u 

in 

U 

O 

•i-j 

a, 

rt 

be 

'S 

"d o 


"d 

fl 


u 

-4-1 

d 
o 

^ o 

s ^ 

a gvj 
V 

a .a 


i/i 

C 

O 

tn 

V 


V ^ 

C^-d 

xi <u 
<u ^ 

4J ds 
<j 

flj "d 

^ a 

xi ” 

rH 

w o 
c/5 S 

+ 1 

- U 

^ o 

J ci 

d 


Cm 

o 

Sm 

u 

rO 

a 

d 

fl 


U5 

d 

o 

in 

ci 

<u 


(U 
CJ 
d 

n bo 
JJ d 

T§ 

Td X 

^ .iS 

(U N 

•D ^ 

D G 

in O 

V) 

cj M 

d 

1) d 
u ^ 
d 
cfl 
Td 
d 
d 


I 


d 

a 

t/5 

Td 

d 


r5 C^ 

Ij d 
^ u 

4j 

■i-i 

U5 

CM I 

Ji -B 

-a 3 

,d 0 
H c/5 





o 

a> 

Oi 

o 

X 

X 

X 

i> 

X 

X 

05 

CM 


X 




q 

q 

q 

X 



q 

on 


rH 

X 

rH 

CM 




CM 

i-H 

iH 

rH 

d 

d 

d 

rH 

d 

d 

d 

d 

d 

d 

d 



O 

o 

+1 

+1 

+ 1 

+ 1 

+1 

+ l 

+ l 

+ 1 

+ 1 

+1 

+1 

+1 

+1 

+ 1 


X 

CM 

CM 

o 

TjH 


o 

05 

rH 


CM 



X 

X 

05 

05 



q 

q 

O' 

q 

X 

X 

X 

05 

X 

X 


CM 

X 

on 


II 


»rj 

d 

J> 

CO 

d 

d 

CO 

d 

d 

d 

rH 

d 

d 

d 








CM 


Th 

X 

X 

X 

X 

o- 

O' 

X 




CM 

i> 

i> 

X 

CM 



Tb 



X 


CM 

rH 

c/: 

Q 

rH 

o 

f— s 

rH 

+1 

rH 

+1 

CO 

+1 

d 
+ 1 

d 
+ 1 

o 

d 
+ l 

d 
+ 1 

d 
+ 1 

d 
+ 1 

d 

+1 

d 

+1 

d 

+1 

d 

+1 

u 


CM 

CM 

l-H 

CM 

CM 

X 

X 


rH 

X 


X 

i> 

X 


X 

§ 



q 

05 

05 

q 

T 


q 

q 

X 


q 

CM 

X 


hJ 



irj 

X 

d 

CO 

d 


CM 


d 

d 

rH 

d 

d 

d 

k-] 





rH 












W 

H 



O) 

O) 

on 

X 

X 

X 

05 

X 

X 

O' 

X 

X 

X 





q 

m 

X 

CM 

X 

q 

X 

CM 

rH 

X 

o 

CM 




rH 

d 

lH 

d 

d 

d 

d 

1 — 1 

d 

d 

d 

d 

d 

d 




o 

o 

+1 

+ 1 

+1 

+ 1 

+ 1 

+ l 

+ 1 

+ 1 

+ 1 

+ 1 

+ 1 

+ 1 

+ 1 

o 


X 

CM 

CO 

on 

X 

X 

CM 

o 

X 

X 

X 

O' 

o 

X 

X 





X 

on 

q 

q 


X 

q 

q 

q 


X 

o 

X 



II 


CM 

d 

rH 

CO 

d 

rH 

CM 

d 

d 

d 

rH 

d 

d 






















CO 

CM 

X 

o 

iH 

O' 


X 

X 


rH 

X 

X 



H 


X 

I— H 

X 

X 

tH 

iH 

O' 

iH 

CM 



o 

CM 




o 

d 

rH 

d 

d 

d 

d 

d 

d 

d 


d 

d 

d 




o 

o 

+ 1 

+ 1 

+ 1 

+1 

+1 

+l 

+1 

+1 

+1 

o 

+ 1 

+1 

+ 1 

o 



CM 

UO 

X 

o 

X 



CM 

X 

o 


X 

X 

CM 





1> 

in 

q 

X 

iH 

iH 

q 

CM 

X 


q 

o 










d 

d 

H 

d 

d 


rH 

d 

d 





00 

CM 

X 

o 

X 

O' 

rH 

X 

X 

o 

X 

X 

X 

CM 




i> 

o> 

I — 1 

X 

X 

X 

iH 

CM 

X 


rH 

rH 

fH 

CM 



CM 

o 

l-H 

d 

rH 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 



+1 

+ 1 

+ 1 

+ 1 

+l 

+ l 

+1 

+ 1 

+1 

+ 1 

+1 

+ 1 

+1 

+1 


,,»— V 

o 
















X 

CM 


o 

X 


X 

05 

CM 

o 



iH 

X 

05 

X 


CM 


o 

o 

X 


05 

X 

CM 

q 

rH 


Tb 

O' 

tH 



II 



i> 

d 

CO 

CM 

d 

d 




d 

d 

d 

d 



rH 

















o 

X 

j> 

X 

X 

05 

o 

J> 

X 


o 

CM 


CM 

CO 

0 



q 

q 

rH 

iq 

X 

CM 

rH 

CM 

rH 

q 

H 

rH 

o 

rH 

P 

rH 

O 

(*»S 

CM 

+ 1 

d 

+1 

rH 

+ 1 

d 

+ 1 

d 

+l 

d 

+l 

d 

+1 

d 

+ 1 

d 

+1 

d 

+1 

o 

+1 

d 

+1 

d 

+1 

d 

+ 1 



GNT 

X 

CM 

O) 

T— 1 

-O 

X 

CM 

O' 

rH 

05 

X 

o 


o 

»-) 



CM 

CM 



q 

q 

q 

q 

rt< 

X 

q 

q 

q 

q 

|Z 



CM 

in 

d 

CM 

CM 

d 

d 

d 

d 

d 

d 

d 

d 

d 

s 



rH 














M 

r . 

















ce 



X 

o 

-t^ 

X 

X 

O' 

05 

-O' 

05 

TfH 

05 

X 

05 

X 




on 


05 

X 

X 

q 

o 


X 

CM 


iH 

iH 

CM 

* 


Xi 

CM 

lH 

d 

d 

d 

rH 

d 

d 

d 

d 

d 

d 

d 

d 

H 


o 

+ 1 

+ 1 

+1 

+1 

+ l 

+ l 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

P 

0 

<Xt 

X 

O 

c^r 

X 

o 

o 

o 

X 

on 

X 

X 

X 

X 

CO 

O' 

O' 

X 

1-H 

q 

X 

X 

X 

on 

O' 

rH 

q 

rH 

X 

CO 

CM 

CM 

O' 

II 


rH 

rH 

in 


CM 

CM 

CO 

d 

rH 

rH 

d 

d 

d 

d 

d 




X 

o 

X 

05 

rH 

X 


X 

X 


CO 

rH 

.!> 

cr5 




X 

q 

in 

O' 


q 


q 

CM 


H 

CM 

O 

X 




CM 

rH 

d 

d 

d 

CM 


d 

d 


d 

d 

d 

d 



o 

o 

,»^ 

+ 1 

+ 1 

+ 1 

+1 

+ l 

+ l 

o 

+1 

+ 1 

o 

+ 1 

+ 1 

+ 1 

+ 1 




r> 


i> 

t'» 

X 

O 


o 

X 


CO 

X 

O' 





CM 

q 


q 

rH 




O' 


CO 

X 

o 

X 




1-H 

in 

CM 

CM 

rH 

Tb 


H 

d 


d 

d 

d 

d 




tH 















IZ) 

M 

M 

U 

bj 

Pl, 

V5 


U 


P ^ ^ 

S i 3 

« P u 



d 

o 


o 

u 

U 


d 

o 

a 

a 

o 

u 


d 

o 

a 

a 

o 

o 

d 

p5 


CM 


Cl 

O 


O 

u 

C 

P 

m 


cl 

o 

e 

a 

o 

o 

X 

(D 

M 

o 

u 

-4-1 

u 

X 


d 

X 

•g 

X 

c 


VD X 
CM X 

I 2 
O o 

d 

II A 

§ 5^ 
e d 
a ^ 

o y 
u X 

a 

1> c" 

. . w 

^ o 

^ C 

Ij 1) 
u 

c 

rfl 
X 
d 
d 
X 
cfl 

OJ 
> 


d 

o 

'C 

d 

Oh 

a 

0 

u 


d 

2 ^ 


423 


424 


Jensen et al. 


VoL. 39, No. 4 



Figure 3. Raptor species numbers and diversity in the 
Southwestern and Western Llanos of Venezuela for com- 
bined years during the wet season (2000 and 2001) and 
the dry season (2001 and 2002). 



Number of Routes Surveyed 


Figure 4. Species accumulation graphs for raptors de- 
tected in the Southwestern and Western Llanos of Ven- 
ezuela for combined years during the wet seasons (2000 
and 2001) and the dry seasons (2001 and 2002). 


Overall, the Southwestern and Western regions 
of the Llanos lacked the eagle diversity character- 
istic of African savannas (Thiollay 1978). Nonethe- 
less, these Llanos regions supported numbers of 
raptor genera (21) similar to those found in the 
seasonally-flooded savannas and agricultural-forest 


mosaic habitats of Kidepo Valley National Park of 
Uganda (22; Thiollay 1978). Consequently, migra- 
tory species in the Southwestern and Western Lla- 
nos only comprise a small portion of the raptor 
community (7%), whereas migratory species ac- 
count for 33% of the raptor community in Kidepo 
Valley National Park of Uganda. 


Table 3. Relative abundance of species and mean number of individuals (± SE) detected per route (mean/route) 
during raptor surveys for combined years in the wet (2000 and 2001) and the dry season (2001 and 2002) in the 
Southwestern and Western Llanos of Venezuela. 


Relative 

Abundance 

Class* 


Southwestern (N = 28) Western {N = 22) 

Species Status*’ Wet'^ DrV Wet^ DrV' 


Very common 

Crested Caracara 

R 

10.21 

+ 

1.37^ 

13.80 

H- 

1.88 

2.98 

± 0.88^ 

6.00 

+ 

1.62 

Yellow-headed Caracara 

R 

6.09 


0.70 

6.70 

+ 

0.74 

5.21 

± 0.97 

8.05 

H- 

1.25 


Savanna Hawk 

R 

4.45 

- 1 - 

0.63 

6.34 


1.06 

3.16 

± 0.76 

12.70 


3.92 


Roadside Hawk 

R 

3.25 


0.62 

3.59 


0.76 

5.18 

± 0.81 

5.41 


0.80 

Common 

Black-collared Hawk 

R 

2.36 


0.53 

2.73 


0.55“ 

0.52 

± 0.21 

0.66 


0.20“ 


Snail Kite 

R 

3.00 


0.94 

0.93 


0.35 

0.73 

± 0.22 

0.23 


0.09 


American Kestrel 

R/NAM 

0.25 

+ 

0.13 

0.23 

+ 

0.10“ 

2.21 

± 0.76 

2.37 


0.57“ 


White-tailed Hawk 

R 

1.41 

+ 

0.24 

0.82 

+ 

0.18 

0.48 

± 0.24 

0.68 


0.17 

Uncommon 

Aplomado Falcon 

R 

0.61 

+ 

0.17 

0.84 

+ 

0.20 

0.57 

± 0.13 

0.60 


0.16 


Great Black Hawk 

R 

0.84 

+ 

0.30 

1.21 

+ 

0.37 

0.59 

± 0.19 

0.43 

+ 

0.18 


White-tailed Kite 

R 

0.32 

+ 

0.09 

0.39 


0.12 

1.48 

± 0.40 

1.50 


0.32 


Crane Hawk 

R 

0.30 


0.10 

0.61 


0.14 

0.27 

± 0.11 

0.55 


0.19 


Laughing Falcon 

R 

0.23 

-h 

0.12 

0.18 


0.13 

0.93 

± 0.27 

1.09 

-4- 

0.38 


Osprey 

NAM 

0.93 


0.2U 

0.38 


0.14 


O'* 

0.28 

+ 

0.11 


“Relative abundance class: Very common = 10-26% of zill individuals detected; Common = 3-5%, Uncommon = 1-2%. 

*^ Status: R = resident species, NAM = North American migratory species, R/NAM = resident and North American migratory 
population. 

^ Means denoted by “d” differed (P < 0.0018) between regions for the wet season. Means denoted by “D” differed {P < 0.0018) 
between regions for the dry season. All other comparisons were not significant (P > 0.0018). 



December 2005 


The Raptor Community of the Li^nos 


425 


Table 4. Rare species detected during raptor surveys for combined years in the wet (2000 and 2001) and the dry 
season (2001 and 2002) in the Southwestern and Western Llanos of Venezuela. 


Relative 



Southwestern {N = 28) 

Western {N 

= 22) 

Abundance 







Class'^ 

Species 

Status'’ 

Wet 

Dry 

Wet 

Dry 

Rare 

King Vulture 

R 

X 


X 

X 


Harris’s Hawk 

R 

X 

X 

X 

X 


Gray-lined Hawk 

R 

X 

X 

X 

X 


Zone-tailed Hawk 

R 

X 

X 

X 

X 


Slender-billed Kite 

R 

X 


X 



Bat Falcon 

R 

X 

X 

X 

X 


Plumbeous Kite 

R 




X 


Hook-billed Kite 

R 

X 

X 

X 

X 


Long-winged Harrier 

R 

X 

X 


X 


Common Black Hawk 

R 

X 





Peregrine Falcon 

NAM 


X 


X 


Short-tailed Hawk 

R 

X 


X 

X 


Grey-headed Kite 

R 


X 


X 


Pearl Kite 

R 

X 


X 



“Relative abundance class: Rare = <1% of all individuals detected. 

’’ Status: R = nonmigratory populations, NAM = North American migratory population. 


Regional and Seasonal Patterns. The Western 
Llanos was characterized by higher levels of raptor 
diversity. This region underwent a period of exten- 
sive deforestation prior to 1825, followed by forest 
regeneration through 1950, and another period of 
deforestation by 1975 (Veillon 1976). Forest ex- 
ploitation cycles, coupled with agricultural activity 
and year-round water sources, have resulted in a 
dynamic mosaic of forest, savanna, agricultural, 
pasture, and early successional habitats that likely 
account for the high raptor diversity in this region. 

Overall, raptor communities in the Southwest- 
ern and Western Llanos were similar. However, 
varying vegetation cover types, large-scale flooding, 
and the availability of year round water sources 
during the dry season almost certainly influence 
the raptor community. For example, Balgooyen 
(1989) reported the American Kestrel preferred 
the forest-agriculture mosaic habitats in the Llanos. 
Our data indicated this pattern was pronounced in 
the dry season, when American Kestrel numbers 
increased in the Western Llanos, likely due to the 
arrival of wintering North American migrants. Fur- 
thermore, several examples of seasonal influences 
are apparent in the Southwestern Llanos, where 
flooding of savanna and gallery forest is extensive 
in the wet season, but also where wetland com- 
plexes persist throughout the year. The higher 
numbers of the Black-collared Hawk in this region 


during the dry season were likely explained by the 
presence of year-round wetland complexes. Simi- 
larly, the higher numbers of Osprey in the South- 
western Llanos versus Western Llanos during the 
wet season indicated that Osprey used both regions 
during the dry season, but first-year birds spent the 
wet season in the Southwestern Llanos, where 
flooding was extensive. Therefore, the year-round 
availability of wetland complexes and the extensive 
inundation of savanna in the wet season likely ex- 
plains why these aquatic-dependent species are 
more abundant in this region. 

Raptor Distribution. Our results on the relative 
abundance and distribution of the 14 most com- 
mon species in our study were consistent with pre- 
vious findings. However, 4 of 14 less common spe- 
cies were not expected to occur in the Llanos, or 
there was little information on their distribution 
and seasonal occurrence: Common Black Hawk (5 
individuals). Pearl Fate (6), Plumbeous Kite (9), 
and Short-tailed Hawk (5). The Pearl Kite was re- 
ported as scarce or absent in the Llanos (Hilty 
2003), but was detected in greater numbers than 
many species considered uncommon residents. Al- 
though the distribution of the Short-tailed Hawk 
was previously unknown in the Llanos (Hilty 2003), 
we detected this species in low numbers through- 
out the study area. Our observation that the Plum- 
beous Kite was absent during the wet season sug- 


426 


Jensen et al. 


VoL. 39, No. 4 


gests the Llanos population was similar to those of 
Central America, Mexico, and Trinidad. All of 
these raptors migrate southward in August and 
September and return north to breed in February 
and March (Ferguson-Lees and Christie 2001). 

Conservation Implications. The rich raptor com- 
munity of the Southwestern and Western Llanos is 
comprised of wetland-dependent and upland-ter- 
restrial species, both nonmigratory and migratory, 
of which several appear seasonally nomadic and 
may depend on both regions as they move in and 
out in response to the wet and dry seasons. This 
pattern suggests that, at least in part, the diverse 
raptor community in the region owes its origins to 
a combination of a large landscape with a substan- 
tial seasonal influx of water and the forest-agricul- 
tural mosaic that creates a temporally and spatially 
diverse mix of habitats. 

The Harris’s Hawk {Parabuteo unicinctus ) , Savan- 
na Hawk, and White-tailed Hawk are now absent 
from the Cauca Valley in Columbia as a result of 
landscape homogenization and wetland depletion 
(Alvarez-Lopez and Kattan 1995). We commonly 
detected the Savanna Hawk and White-tailed 
Hawk, and to a lesser extent the Harris’s Hawk, 
throughout the Southwestern and Western Llanos, 
which suggests large-scale landscape homogeniza- 
tion and wetland degradation are not yet occurring 
in these regions. In contrast, because hawk-eagles 
may be less tolerant of human alterations on the 
landscape than many other raptors (Burnham et 
al. 1994), the Ornate Hawk-Eagle may have been 
affected by historical deforestation and subsequent 
lack of suitable habitats in these regions. However, 
with the exception of the Ornate Hawk-Eagle, the 
raptor community in the Llanos may represent a 
community model for Neotropical savanna-forest- 
agricultural regions. 

Importantly, additional studies on the raptor 
community in the Llanos are required to better 
understand abundance patterns and seasonal fluc- 
tuations of raptors. Eurthermore, research is need- 
ed to evaluate local areas where deforestation, in- 
tensive agriculture, and man-made impoundments 
are currently expanding, which may be useful for 
assessing the long-term stability of the current rap- 
tor community. Finally, comparisons between the 
Llanos and other Neotropical wetland-savanna 
complexes, such as the Brazilian Pantanal, will help 
determine the scope and representation of our 
findings. 


Acknowledgments 

Duck’s Unlimited and the U.S. Department of Agri- 
culture Forest Service provided the major financial sup- 
port for this project, along with support from the Hawk 
Mountain Sanctuary. Special thanks to our field assistants 
Alexis Araujo, Mireya Barrera, David Peraza, Mariana Es- 
cobar, Maria Doris Escovar, Andres DeGraf, Sara Seijas, 
and Graciela Barrera and to the Universidad Nacional 
Experimental de Los Llanos Occidentales Ezequiel Za- 
mora (UNELLEZ) in Guanare for their collaboration 
with the project and the generous use of their facilities. 
Thanks to Don Taphorn, Jose Angel Anez of AsoMuseo, 
Luis Altuve, Jose Gregorio Quintero, Gilberto Rios, An- 
tonio Utrera, Andres Seijas, Franklin Rojas-Suarez of 
Pro Vita, Alvaro Velasco from Profauna in Caracas, Jose 
Gregorio Garcia Tenia and Yuri Cedeno from Profauna 
in San Fernando, Richard Schargel and Heberto Pacheco 
from BioCentro, Clemencia Rodner and Robin Restall 
from the Audubon Society in Caracas, and Luis Gonzalez 
Morales from the Universidad Central de Venezuela. Fi- 
nally we gratefully thank the following landowners in the 
Llanos for access to their properties and use of facilities, 
the Maldonado family and staff members Alexis Aguirre, 
Jose Ayarzaguena, and Sara Candela of Hato El Frio; Ing. 
Jesus Pacheco and Enrique Loreto who authorized visits 
to Hato El Cedral, where staff members Edgar Chiapan- 
na and Ing. Tulio Aquilera provided assistance; the Bran- 
ger family at Hato Pinero, where Edgar Useche autho- 
rized visits; staff at Hato Fernando Corrales; and Don 
Porfirio Martinez from Mantecal. This is Hawk Mountain 
Sanctuary contribution to conservation science number 
119. 

Literature Cited 

Alverez, E., D.H. Ellis, D.G. Smith, AND C.T. Larue 
1996. Diurnal raptors in the fragmented forest of Si- 
erra Imataca, Venezuela. Pages 263-273 in D. Bird, D. 
Varland, andj. Negro [Eds.], Raptors in human land- 
scapes. Academic Press, London, England. 
A1.VAREZ-L0PEZ, H. AND G.H. Kattan. 1995. Notes on the 
conservation status of resident diurnal raptors of the 
Middle Cauca Valley, Colombia. Bird Cons. Inti. 5:341- 
348. 

Balgooyen, T.G. 1989. Natural history of the American 
Kestrel in Venezuela. J. Raptor Res. 23:85-93. 

Beai., K.G. AND H.J. Khamis. 1991. A problem in statistical 
analysis: simultaneous inference. Condor 93:1023— 
1025. 

Beissinger, S.R., B.T. Thomas, and S.D. Strahi.. 1988. 
Vocalizations, food habits, and nesting biology of the 
Slender-billed Kite with comparisons to the Snail Kite 
Wilson Bull. 100:604—616. 

Bibby, C.J., N.J. Collar, M.J. Crosby, M.F. Heath, C.H. 
Imboden, T.H. Johnson, AJ. Long, A.J. Sattersfield, 
AND S.J. Thirgood. 1992. Putting biodiversity on the 
map: priority areas for global conservation. ICBP, 
Cambridge, England. 

Bierregaard, R.O., Jr. 1998. Conservation status of birds 


December 2005 


The Raptor Community of the Llanos 


427 


of prey in the South American tropics. J. Raptor Res. 
32:19-27. 

Bildstein, K.L., W. Schelsky, J. Zalles, and S. Ellis. 
1998. Conservation status of tropical raptors./. Raptor 
Res. 32:3-18. 

Bulla, L., J. Pacheco, and G. Morales. 1990. Seasonally 
flooded Neotropical savanna closed by dikes. Pages 
177-210 in A.I. Breymeyer [Ed.], Ecosystems of the 
world: managed grasslands. Regional Studies 17A. El- 
sevier Science, Amsterdam, Netherlands. 

Bunn, A.G., W. Klein, and K.L. Bildstein. 1995. Time- 
of-day effects on the numbers and behavior of non- 
breeding raptors seen on roadside surveys in eastern 
Pennsylvania./. Field Ornithol. 66:544—552. 

Burnham, W.A., D.F. Whitacre, and J.P. Jenny. 1994. The 
Maya Project: use of raptors as tools for conservation 
and ecological monitoring of biological diversity. Pag- 
es 257-264 in B.-U. Meyburg and R.D. Chancellor 
[Eds.], Raptor conservation today. WWGBP/The Pica 
Press, Tonbridge, Kent, England. 

Cole, M.M. 1986. The savannas: biogeography and geo- 
botany. The savanna woodlands, savanna grasslands 
and low tree and shrub savannas of northern tropical 
America. Academic Press, London, England. 

Colwell, R.K. 2000. Estimates: statistical estimation of 
species richness and shared species from samples, ver- 
sion 6. Obi. University of Connecticut, Storrs, CT, 
U.S.A. 

Droege, S. 1990. The North American Breeding Bird 
Survey. Pages 1-4 mJ.R. Sauer and S. Droege [Eds.], 
Survey designs and statistical methods for the esti- 
mation of avian population trends. U.S. Fish and 
Wildlife Service, Biological Report 90. 

Ellis, D.H., R.L. Glinski, and D.G. Smith. 1990. Raptor 
road surveys in South America. /. Raptor Res. 24:98- 
106. 

Ferguson-Lees, j. and D.A. Christie. 2001. Raptors of 
the world. Houghton Mifflin Company, New York, NY 
U.S.A. 

Fuller, M.R., S.W. Seegar, and L.S. Schueck. 1998. 
Routes and travel rates of migrating Peregrine Fal- 
cons Falco peregrinus and Swainson’s Hawk Buteo swain- 
soni in the western hemisphere. /. Avian Biol. 29:433- 
440. 

Hayek, L.C. and M.A. Buzas. 1996. Surveying natural 
populations. Columbia University Press, New York, NY 
U.S.A. 

Hilty, S.L. 2003. Birds of Venezuela, 2nd Ed. Princeton 
University Press. Princeton, NJ U.S.A. 

Huber, O. and C. Alcaron. 1988. Mapa de vegetacion 
de Venezuela. Minestro del ambiente y de los Recur- 
sos Naturales Renovables. DGHA, Division de Vege- 
tacion, Caracas, Venezuela. 

Jensen, W.J. 2003. The abundance and distribution of fal- 
coniformes in the central and western llanos of Ven- 
ezuela. M.S. thesis. State Univ. of New York, College 


of Environmental Science and Forestry, Syracuse, NY 
U.S.A. 

Kirk, D.A. and J.P. Currall. 1994. Habitat associations 
of migrant and resident vultures in central Venezuela 
/. Avian Biol. 25:327—337. 

Mader, W.J. 1981. Notes on nesting raptors in the Llanos 
Condor 83:48-51. 

. 1982. Ecology and breeding habits of the Savan- 
na Hawk in the Llanos of Venezuela. Condor 84:261- 
271. 

Magurran, A.E. 1988. Ecological diversity and its mea- 
surement. Princeton University Press, Princeton, NJ 
U.S.A. 

Manosa, S., E. Mateos, and V. Pedrocchi. 2003. Abun- 
dance of soaring raptors in the Brazilian Atlantic rain- 
forest. /. Raptor Res. 37:19-30. 

and V. Pedrocchi. 1997. A raptor survey in the 

Brazilian Atlantic rainforest./. Raptor Res. 31:203-207. 

Martell, M.S., C.J. Henny, RE. Nye, and MJ. Solensky 
2001. Fall migration routes, timing, and wintering 
sites of North American Ospreys as determined by sat- 
ellite telemetry. Condor 103:715—724. 

Millsap, B.A. and M.N. LeFranc, Jr. 1988. Road transect 
counts for raptors, how reliable are they? /. Raptor Res 
22:8-16. 

Mittermeier, R.A and C.G. Mittermeier. 1997. Venezue- 
la. Pages 448 — 467 in Megadiversity-Earth’s biologi- 
cally wealthiest nations. Conservation International 
CEMEX de SV, The University of Chicago Press, Chi- 
cago, IL U.S.A. 

, , P.R. Gil, J. Pilgrim, G. Fonesca, T. 

Brooks, and W.R. Konstant. 2003. The llanos. Pages 
265—270 in Wilderness: earth’s last wild places. Con- 
servation International CEMEX de SV, The University 
of Chicago Press, Chicago, IL U.S.A. 

Myers, N., R.A. Mittermeier, C.G Mittermeier, G.A.B. 
DE Fonesca, and J. Kent. 2000. Biodiversity hotspots 
for conservation priorities. Nature 403:853-858. 

Seavy, N.E. and C.K. Apodaca. 2002. Raptor abundance 
and habitat use in a highly-disturbed-forest landscape 
in Western Uganda. / Raptor Res. 36:51-57. 

Silva, J.F and A. Moreno. 1993. Land use in Venezuela 
Pages 239—257 mM.D. Young and O.T. Solbrig [Eds ], 
The world’s savannas; economic driving forces, eco- 
logical constraints, and policy options for sustainable 
land use. UNESCO, Paris, France. 

SoRLEY, C.S. AND D.E. Andersen. 1994. Raptor abun- 
dance in south-central Kenya in relation to land use 
patterns. Afr. J. Ecol. 32:30-38. 

Thiollay, J.-M. 1978. Population structure and seasonal 
fluctuations of the falconiformes in Uganda National 
Parks. East Afr. Wildl. J. 16:145-151. 

. 1984. Raptor community structure of a primary 

rainforest in French Guiana and effect of human 
hunting pressure. Raptor Res. 18:117-122. 

. 1989. Area requirements for the conservation of 


428 


Jensen et al. 


VoL. 39, No. 4 


rain forest raptors and game birds in French Guiana. 
Conserv. Biol. 2:128—137. 

Troth, R.G. 1979. Vegetational types on a ranch in the 
central llanos of Venezuela. Pages 17-30 mJ.F. Eisen- 
berg [Ed.], Vertebrate ecology in the neotropics. 
Smithsonian Institute Press, Washington, DC U.S.A. 

Veillon, J.P. 1976. Las Deforestaciones en la Region de 
los Llanos Occidentales de Venezuela, desde 1950 has- 
ta 1975. Pages 67-112 mL.S. Hamilton, J. Steyermark, 
J.P. Veillon, and E. Mondolfi, [Eds.], Conservacion de 
los Bosques Humedos de Venezuela. Sierra Club y 
Consejo del Bienestar Rural, Caracas, Venezuela. 


Whitacre, D.F. and C.W. Turley. 1990. Further compar- 
isons of tropical rainforest census techniques. Pages 
71-92 mW.A. Burnham, D.F. Whitacre, and J.P. Jenny 
[Eds.], Maya Project; use of raptors as environmental 
indices for design and management of protected ar- 
eas and for building local capacity for conservation in 
Latin America. The Peregrine Fund, Boise, ID U.S.A. 

WOFFINDEN, N.D. AND J.R. MuRPHY. 1977. A roadside rap- 
tor census in the eastern Great Basin, 1973-1974. Rap- 
tor Res. 11:62-66. 

Received 27 May 2004; accepted 15 August 2005 


J. Raptor Res. 39(4):429-438 
© 2005 The Raptor Research Foundation, Inc. 


A COMPARISON OF BREEDING SEASON FOOD HABITS OF 
BURROWING OWLS NESTING IN AGRICULTURAL AND 
NONAGRICULTURAL HABITAT IN IDAHO 

Colleen E. Moulton,^ Ryan S. Brady, and James R. Belthoff 

Department of Biology and Raptor Research Center, Boise State University, Boise, ID 83725 U.S.A. 

Abstract. Through analysis of regurgitated pellets and prey remains collected at nests between 2001- 

02, we characterized diet composition of western Burrowing Owls {Athene cunicularia hypugaea) in the 
Snake River Birds of Prey National Conservation Area (NCA) of southwestern Idaho. We hypothesized 
that diet differs between owls nesting in agricultural and nonagricultural habitat, because at least one 
important prey species, montane voles (Microtus montanus), occurs predominately in the former. From 
859 pellets, we identified 7402 prey items representing 23 species, and identified 403 prey remains of 
19 species. Invertebrates dominated the diet in numbers of prey, whereas rodents contributed the great- 
est biomass. Montane voles, which were not present in pellets in nonagricultural areas, represented the 
greatest percent biomass of pellets in agricultural areas. Invertebrates (predominately Gryllidae) also 
were more abundant in diets of owls nesting in agricultural habitat. Pellets of owls nesting in agricultural 
areas had greater species richness, whereas pellets from nonagricultural areas had greater species even- 
ness and broader food-niche breadths. Finally, we estimated food-niche breadth of Burrowing Owls in 
the NCA to be broader than previously reported. 

Key Words: Burrowing Owl; Athene cunicularia; agriculture; food habits; food-niche; Idaho. 


UNA COMPARACION DE LOS hAbITOS ALIMENTICIOS DE INDIVIDUOS NIDIFICANTES DE 
ATHENE CUNICULARIA EN AMBIENTES AGRICOLAS Y NO AGRICOLAS EN IDAHO 

Resumen. — A traves del analisis de egagropilas y de restos de presas recolectados en nidos en 2001 y 
2002, caracterizamos la composicion de la dieta de Athene cunicularia hypugaea en el Area Nacional de 
Conservacion de Aves de Presa Snake River, sudoeste de Idaho. Nos planteamos la hipotesis de que la 
dieta difiere entre las lechuzas que nidifican en ambientes agricolas y no agricolas, debido a que al 
menos una de las especies de presa importantes, Microtus montarius, se encuentra predominantemente 
en las areas agricolas. De un total de 859 egagropilas, identificamos 7402 items de presas correspon- 
dientes a 23 especies, e identificamos 403 restos de presas provenientes de 19 especies. Los invertebrados 
dominaron la dieta en terminos del numero de presas, mientras que los roedores representaron la 
mayor biomasa. Microtus montanus no estuvo presente en las egagropilas de las areas no agricolas y 
represento el mayor porcentaje de biomasa en las egagropilas de las areas agricolas. Los invertebrados 
(predominantemente Gryllidae) tambien fueron abundantes en las dietas de las lechuzas que nidifica- 
ron en los ambientes agricolas. Las egagropilas de las lechuzas que nidificaron en las areas agricolas 
presentaron mayor riqueza de especies, mientras que las provenientes de las areas no agricolas presen- 
taron mayor equidad y nichos alimenticios mas amplios. Finalmente, estimamos que el nicho alimenticio 
de A. c. hypugaea en el area silvestre de conservacion estudiada es mas amplio de lo que habia sido 
informado previamente. 

[Traduccion del equipo editorial] 


Agricultural practices historically have provided 
many different types of wildlife habitat, including 
shelterbelts, hedgerows and fencerows, cultivated 
fields, and fields in rotation. Although some spe- 
cies nest, seek cover, and forage in these habitats, 


1 Present address: Wildlife Bureau, Idaho Department of 
Fish and Game, P.O. Box 25, Boise, ID 83707 U.S.A.; 
email address: cmoulton@idfg.idaho.gov 


many wildlife populations have declined signifi- 
cantly in areas of agricultural conversion (Carlson 
1985, Murphy 2003). In fact, there is mounting ev- 
idence that converting natural landscapes into ag- 
ricultural use can affect a wide array of wildlife 
populations through erosion, exposure to herbi- 
cides and pesticides, and destruction of nesting 
and cover habitat (Carlson 1985, Jahn and Schenck 
1991, Gervais et al. 2000). These effects may be 


429 


430 


Moulton et al. 


VoL. 39, No. 4 


amplified by the shift from small-scale farming 
practices to large-scale monoculture farming seen 
throughout the United States and Canada (Peter- 
john 2003). 

Western Burrowing Owls {Athene cunicularia 
hypugaea) are listed as Endangered in Canada and 
several western U.S. states, and their populations 
are declining in many areas (e.g., James and Espie 
1997, Clayton and Schmutz 1999, Klute et al. 
2003). These owls suffer deleterious effects from 
agricultural practices (James and Fox 1987, Haug 
et al. 1993, Bellocq 1997, Gervais et al. 2000) and, 
in Canada, often avoid agricultural fields (Haug 
and Oliphant 1990, Clayton and Schmutz 1999). 
However, throughout some portions of their west- 
ern U.S. range. Burrowing Owls associate with ag- 
riculture (Rich 1986, DeSante et al. 2004, Moulton 
et al., in press), and they are the only raptor spe- 
cies that shows significant affinity for agriculture in 
southern Idaho (Leptich 1994). Rich (1986) sug- 
gested that proximity to montane voles {Microtus 
montanus) in farmlands could explain some of this 
habitat selection. Moulton et al. (in press) con- 
firmed that owls did not nest in agricultural areas 
because of decreased nest predation or increased 
availability of nesting sites but noted that prey con- 
sumption was greater in agricultural areas. 

If Burrowing Owl nesting distributions can be 
affected by prey, as Rich (1986) and Moulton et al. 
(in press) hypothesize, then diet composition may 
differ for owls occupying agricultural and nonag- 
ricultural areas. Thus, the objective of our study 
was to examine breeding season food habits of 
Burrowing Owls in the Snake River Birds of Prey 
National Conservation Area (NCA), where Burrow- 
ing Owls inhabit both agricultural and nonagricul- 
tural areas. Specifically, we tested the hypotheses 
that (1) diets of owls in agricultural areas contain 
more montane voles than those in nonagricultural 
habitats and (2) because of influences of agricul- 
tural practices, diet diversity and food-niche 
breadths differ. We predicted that Burrowing Owls 
nesting in agricultural habitats would have greater 
prey diversity and broader food-niche breadths 
than owls nesting in nonagricultural habitats. Fi- 
nally, we compared our food-niche breadth esti- 
mates with those of a previous study (Marti et al. 
1993) on raptor food habits in the NCA. 

Methods 

We studied Burrowing Owls nesting within and near 
the NCA in southwestern Idaho during 2001-02. This 
area was once representative of a typical shrub-steppe 
community dominated by large expanses of big sage- 


brush {Artemesia tridentata xvyomingensis; Hironaka et al. 
1983) and other shrubs, and scattered perennial bunch- 
grasses. However, disturbances, such as range fires, mili- 
tary training, grazing, and off-road vehicle use, have 
helped convert much of the area to exotic annual grass- 
lands dominated by cheatgrass {Bromus tectorum), tumble 
mustard {Sisymbrium altissimum), and other non-native 
species (Hironaka et al. 1983). Surrounding areas also 
contained scattered residential homes, paved and dirt 
roads, a military training area, and public lands managed 
by the Bureau of Land Management. Cattle and sheep 
graze much of the area, especially during winter. Irrigat- 
ed agricultural fields (primarily alfalfa, sugar beets, and 
mint) constituted <5% of the NCA and were located pri- 
marily along its margins (USDI 1996). For the purpose 
of this study, we considered Burrowing Owl nests that 
were within 1 km of an irrigated agricultural field, to be 
in “agricultural” habitat (hereafter agricultural nests). 
Agricultural nests were located in the natural vegetation 
surrounding agriculture fields rather than in the irrigat- 
ed portions where crops grew. However, adult owls fre- 
quently hunted within these fields and perched on fence 
posts adjacent to them (Moulton et al. in press). “Non- 
agricultural” habitat was the term we used to categorize 
nests that were greater than 3 km from irrigated fields 
(hereafter nonagricultural nests). Because this distance 
exceeded the typical foraging range of Burrowing Owls 
(Haug and Oliphant 1990, Rosenberg and Haley 2004), 
we are almost certain that owls from nonagricultural 
nests were not collecting prey from or near irrigated 
fields. Nonagricultural areas were generally disturbed 
shrublands and grasslands much like that in the agricul- 
tural areas, but there were no crops or irrigation nearby 

Diet Composition. Regurgitated pellets are reliable in- 
dicators of the diet of Burrowing Owls (Marti 1974), al- 
though amphibians and reptiles can be underrepresent- 
ed in pellets (Thomsen 1971, Haug 1985). Similarly, prey 
remains alone do not provide reliable information re- 
garding overall diet composition, as many prey items con- 
sumed by Burrowing Owls are too small to cache (such 
as small insects) . But, remains provide better information 
than pellets concerning amphibians and reptiles in the 
diet. Therefore, to determine diet composition, we both 
documented prey remains at nests and collected and an- 
alyzed regurgitated pellets. 

Pellet Collection and Analysis. We collected regurgitat- 
ed pellets from tunnel entrances, perches, and nearby 
mounds within 20 m of nest burrows every 3-10 d from 
hatching through 25 d post-hatch (May-June) . For nests 
at which we collected more than 20 pellets (29 of 51 
nests; 22 agricultural, 7 nonagricultural), we analyzed a 
random sample of 20 pellets per nest. For all other nests, 
we analyzed all collected pellets (11.2 — 1.0 [SE] pellets 
per nest, range = 4-19). 

We analyzed and quantified remains of each pellet us- 
ing standard procedures (Marti 1987) and by comparing 
prey species to a museum collection at Boise State Uni- 
versity. Skulls, jaws, dentition patterns, head capsules, 
pronota, elytra, legs, scales and other distinguishing body 
parts helped identify prey. 

Prey Remains. Owls in this study nested in artificial 
burrows deployed for other studies (Smith and Belthoff 
2001, Belthoff and Smith 2003, Brady 2004, Moulton et 
al. in press), which provided access to nest chambers. 


December 2005 


Burrowing Owl Food Habits in Idaho 


431 


where we found most prey remains. We could therefore 
document cached and uneaten prey remains at all oc- 
cupied nests and adjacent satellite burrows (non-nest 
burrows used by owls for roosting, cover, and caching 
prey) . We quantified prey remains each time we excavat- 
ed an artificial burrow (2-5 visits per nest) between 
hatching and 25 d post-hatch. 

Biomass Estimation. We determined biomass of rep- 
resentative mammalian, avian, and amphibian prey using 
Smith and Murphy (1973) and Steenhof (1983). Biomass 
of invertebrate prey species was determined using our 
own estimates obtained from captured live specimens 
(Moulton 2003) and values reported in Smith and Mur- 
phy (1973) and Olenick (1990). 

Statistical Analysis. Because we obtained prey remains 
from only two nonagricultural nests, we did not include 
data from prey remains in diversity calculations described 
below or in statistical comparisons; instead, pellet data 
provided all information used for these calculations and 
comparisons. We determined each prey type as a percent 
of total prey items per nest (percent number) and per- 
cent biomass per nest. 

We determined food-niche breadth for agricultural 
and nonagricultural nests by calculating the reciprocal of 
Simpson’s index (Simpson 1949). We calculated dietary 
evenness using the Alatalo (1981) modification of Hill’s 
(1973) index: F = - l)/(Wi - 1). 

To determine if differences in diets existed between 
owls nesting in each habitat, we compared percent num- 
ber and percent biomass of each prey taxa (vertebrates) , 
class (invertebrates), or order (invertebrates) per nest us- 
ing Wilcoxon’s ranked sums tests (Zar 1999). If there 
were differences between habitats in taxa/ order of prey, 
we then compared species (vertebrates) or families (in- 
vertebrates) of that taxa/order. Because we made multi- 
ple comparisons of prey categories, we adjusted alpha lev- 
els using sequential Bonferroni corrections (Rice 1989). 

To determine if diet diversity differed between agri- 
cultural and nonagricultural nests, we compared food- 
niche breadth (Simpson’s index), species richness (num- 
ber of species in the diet), and dietary evenness (Alatalo’s 
index) using Wilcoxon’s ranked sums tests. Statistical 
analyses were performed using JMPIN V.5 (SAS Institute, 
Inc., Cary, NC), and evaluated at an alpha level of 0.05 
unless otherwise noted. Throughout, we present means 
with their standard errors. 

Results 

Pellet Remains. We analyzed 602 regurgitated 
pellets from 34 agricultural nests and 257 pellets 
from 19 nonagricultural nests. From these, we 
identified 7402 prey items representing 23 differ- 
ent prey species. 

Overall pellet composition. Invertebrates were the 
most frequent prey in pellets, representing 93% of 
prey items; however, they represented only 23% of 
biomass (Table 1). Conversely, vertebrates (ro- 
dents, birds, and herpefofauna) comprised 7% of 
prey items, but 77% of biomass. 

Coleopterans (beetles) and Orthopterans (crick- 
ets, grasshoppers) were the most common inver- 


tebrates in pellets, constituting 47% and 32% of 
total prey, respectively (Table 1 ) . Of Coleopterans, 
ground beetles (Carabidae) and darkling beetles 
(Tenebrionidae) were most common (33% and 
22% of Coleopteran prey items, respectively). Or- 
thopteran prey remains were predominately Grylli- 
dae (crickets), which constituted 73% of Orthop- 
teran prey items. 

Rodents were the most common vertebrates in 
pellets and represented 97% of vertebrates detect- 
ed and 73% of overall prey biomass (Table 1). 
Pocket mice {Perognathus parvus) and deer mice 
(Peromyscus maniculatus) were the most abundant 
rodents (37% and 25%, respectively), but montane 
voles represented the greatest biomass (18%). 

Habitat variation. Invertebrates were the most fre- 
quent prey in pellets for both agricultural and non- 
agricultural nests, representing 95% and 90% of 
total prey items, respectively (Table 2). Vertebrate 
prey (mostly rodents) represented the greatest per- 
cent biomass in both agricultural (76%) and non- 
agricultural (79%) nests. 

Coleopterans were the most common inverte- 
brates in both habitats (Table 2). However, Arach- 
nids contributed the greatest biomass (52%) of 
invertebrates in nonagricultural nests, and Orthop- 
terans contributed the greatest biomass (52%) of 
invertebrates in agricultural nests. Of rodent spe- 
cies found in pellets, deer mice and pocket mice 
were most common in agricultural and nonagri- 
cultural nests, respectively. Pocket mice also con- 
tributed the greatest biomass of rodents at nonag- 
ricultural nests, but montane voles contributed the 
greatest biomass of rodents at agricultural nests. 
Only owls at agricultural nests preyed on montane 
voles (Table 2). 

Agricultural and nonagricultural nests did not 
differ in percent biomass of vertebrates or inver- 
tebrates (Table 3) . However, agricultural nests had 
a greater percent number of invertebrates, and 
nonagricultural nests had a greater percent num- 
ber of vertebrates. Pellets from agricultural nests 
had greater percent number and percent biomass 
of montane voles (Table 4). Nonagricultural nests 
had greater percent number and biomass of pock- 
et mice (Table 4) . Among invertebrates. Arachnids 
and Orthopterans differed between habitats (Table 
3). Solpugida (windscorpions) and Acrididae oc- 
curred in greater percent number and biomass in 
pellets of nonagricultural nests, while Gryllidae oc- 
curred in greater number and biomass in pellets 
at agricultural nests (Table 5). 

For all nests combined, food-niche breadth was 


432 


Moulton et al. 


VoL. 39, No. 4 


Table 1. Mean (±SE) percent number and percent biomass per nest of prey items detected in pellets collected at 
53 Burrowing Owl nests in southwestern Idaho, 2001-02. 


Prey Category 

Percent Number 

Percent Biomass 

Mammals 

6.7 

“h 

0.7 

72.9 

— b 

2.5 

Spermophilus mollis 

0.2 


0.1 

10.2 

-1- 

3.4 

Thomomys townsendii 

0.2 


0.1 

12.4 


3.5 

Perognathus parvus 

2.5 

+ 

0.6 

12.5 

+ 

2.5 

Dipodomys ordii 

0.6 

-h 

0.2 

9.4 

+ 

2.4 

Reithrodontomus megalotis 

0.1 

— K 

0.1 

0.3 

+ 

0.1 

Peromyscus maniculatus 

1.7 

“b 

0.4 

11.1 

-b 

2.5 

Mus musculus 

0.2 


0.1 

1.2 

“b 

0.5 

Microtus montanus 

0.9 


0.2 

13.2 

-b 

2.8 

Rodent — unidentified^ 

0.4 

“h 

0.1 

2.7 

-b 

0.7 

Birds — unidentified^ 

0.2 

H- 

0.1 

2.2 


0.8 

Reptiles and Amphibians'^ 

<0.1 

-h 

0.1 

1.9 

-b 

1.2 

Arachnids 

13.8 

-h 

1.9 

6.3 

-b 

1.0 

Scorpionida 

5.8 

-h 

1.0 

3.5 

-b 

0.9 

Solpugida 

8.0 

H- 

1.5 

2.8 

-b 

0.5 

Orthopterans 

31.6 

-b 

3.2 

9.6 

-b 

1.6 

Acrididae 

2.9 

-b 

0.6 

0.7 

fr 

0.1 

Gryllidae 

23.2 

-b 

3.5 

7.7 

-b 

1.7 

Unknown Orthoptera 

5.5 

-b 

0.9 

1.2 

-b 

0.2 

Dermapterans (Forficulidae) 

0.4 

-b 

0.2 

0.2 

H- 

0.1 

Homopterans (Cicadidae) 

0.1 

-b 

0.1 

0.1 


0.1 

Coleopterans 

47.0 

-b 

2.6 

6.8 


0.7 

Carabidae 

15.7 


2.2 

1.9 

-b 

0.3 

Scarabidae 

7.5 


1.3 

1.1 

H- 

0.2 

Silphidae 

8.0 


1.4 

1.1 

H- 

0.2 

Tenebrionidae 

10.6 

-b 

1.9 

2.3 

-H 

0.4 

Coleoptera — unidentified 

5.3 

+ 

1.2 

0.5 

+ 

0.1 

Total vertebrates 

6.9 

H- 

0.7 

77.0 


2.1 

Total invertebrates 

93.1 

+ 

0.7 

23.0 

-b 

2.1 


^ Mouse species: likely P. parvus, R. megalotis, P. maniculatus, or M. musculus. 

^ Likely Eremophila alpestris or Sturnella neglecta. 

Includes Bufo woodhousei, Phrynosoma platyrhinos, and unknown snake species. 


4.22 ± 0.22 {N = 53). Nonagricultural {N = 19) 
nests had greater species evenness than agricultur- 
al {N = 34) nests (0.76 ± 0.03 versus 0.60 ± 0.02; 
Z = 3.89, P < 0.001) and broader food-niche 
breadth (5.21 ± 0.33 versus 3.67 ± 0.25; Z = 3.24, 
P — 0.001). However, agricultural nests had higher 
species richness (11.82 ± 0.40 versus 9.79 ± 0.54; 
Z = -2.69, P = 0.007). 

Prey Remains. We recorded cached and other 
uneaten prey remains at 43 nests {N = 41 agricul- 
tural, N— 2 nonagricultural) and documented 403 
prey items representing 19 species (Table 6). Be- 
cause we had so few nonagricultural nests, we 
made no comparisons between habitats and 
pooled data from all nests for descriptions of prey 
remains. 


Although common in pellets, invertebrate prey 
remains were uncommon in nest burrows {N = 50 
individual invertebrate prey items) . The majority of 
prey remains in both percent number (87.6%) and 
percent biomass (99.7%) were vertebrates, most of 
which were rodents. Of rodent species, montane 
voles were most common by number (36%), and 
pocket gophers represented the greatest biomass 
(50%). 

Although rare in pellets, we occasionally found 
herpetofauna (N — 38) and birds {N = 18) cached 
in burrows. Woodhouse’s toads {Bufo woodhousei) 
were the most common (92%) herpetofauna in 
nest burrows. All toads were in nests adjacent to 
agricultural fields. Burrowing Owl nestlings were 
the most common (50%) cached avian prey item 


December 2005 


Burrowing Owl Food Habits in Idaho 


433 


Table 2. Mean (±SE) percent number and percent biomass per nest of prey items detected in pellets of Burrowing 
Owls nesting in agricultural (N = 34) and nonagricultural (N = 19) habitats of southwestern Idaho, 2001-02. 


Agricultural Nonagricultural 


Prey Percent No. Percent Biomass Percent No. Percent Biomass 


Mammals 

4.9 


0.8 

70.7 

-h 

3.1 

10.1 ± 

1.1 

76.9 ± 4.1 

Spermophilus mollis 

0.1 

-h 

0.1 

5.3 

+ 

4.1 

0.5 ± 

0.2 

0.2 ± 0.1 

Thomomys townsendii 

0.4 

-H 

0.1 

1.9 


4.1 



— 

Perognathus parvus 

0.7 


0.6 

4.9 

± 

2.6 

5.6 ± 

0.8 

26.0 ± 3.4 

Dipodomys ordii 

0.4 

± 

0.2 

4.9 


2.8 

1.2 ± 

0.3 

17.5 ± 3.8 

Peromyscus maniculatus 

1.4 

+ 

0.5 

10.5 


3.1 

2.1 ± 

0.7 

12.3 ± 4.2 

Mus musculus 

0.2 


0.1 

1.9 

+ 

0.6 

— 


— 

Microtus montanus 

1.4 

-h 

0.3 

20.6 

+ 

3.1 

— 


— 

Birds — unidentified^ 

0.1 

H- 

0.1 

2.3 

+ 

1.0 

0.3 ± 

0.1 

2.0 ± 1.3 

Reptiles and Amphibians*’ 

0.1 

-+- 

0.1 

2.8 

+ 

1.5 

0.1 ± 

0.1 

0.3 ± 2.0 

Arachnida 

6.7 


1.7 

3.7 

+ 

1.1 

26.5 ± 

2.3 

10.9 ± 1.4 

Scorpionida 

3.3 

-h 

1.1 

2.0 


1.0 

10.3 ± 

1.5 

6.0 ± 1.4 

Solpugida 

3.4 

H- 

1.6 

1.7 

± 

0.6 

16.2 ± 

2.1 

4.9 ± 0.8 

Orthoptera 

40.0 


3.5 

12.7 

± 

1.9 

16.7 ± 

4.7 

4.0 ± 2.6 

Acrididae 

1.8 

-t- 

0.8 

0.5 

H- 

0.2 

4.9 ± 

1.0 

1.1 ± 0.2 

Gryllidae 

34.8 

± 

3.5 

11.4 

-h 

1.9 

2.3 ± 

4.7 

1.0 ± 2.5 

Coleoptera 

47.6 

± 

3.2 

7.5 


0.8 

46.1 ± 

4.3 

5.7 ± 0.1 

Carabidae 

21.4 

-t- 

2.4 

2.7 


0.4 

5.5 ± 

3.2 

0.7 ± 0.5 

Scarabidae 

5.6 

-h 

1.5 

1.0 


0.3 

10.9 ± 

2.0 

1.3 ± 0.4 

Silphidae 

5.1 


1.7 

0.9 

+ 

0.3 

13.2 ± 

2.3 

' 1.4 ± 0.4 

Tenebrionidae 

11.1 

-h 

2.3 

2.5 

H- 

0.6 

9.6 ± 

3.1 

1.9 ± 0.7 

Total vertebrates 

5.0 

+ 

0.8 

75.8 

+ 

2.6 

10.3 ± 

1.1 

79.1 ± 3.5 

Total invertebrates 

95.0 

+ 

0.8 

24.2 


2.6 

89.7 ± 

1.1 

20.9 ± 3.5 


“ Likely Eremophila alpestris or Sturnella neglecta. 

Includes Bufo woodhousei, Phrynosoma platyrhinos, and unknown snake species. 


we found. These Burrowing Owl nestlings all were 
individuals from nests other than the nest in which 
we found them. Whether they wandered into the 
nest on their own and subsequently starved or were 
killed or were taken directly from their nest is un- 
known. We suspect that adults tending nearby nests 
preyed upon these nestlings because they frequent- 
ly were too young to have wandered into nests oth- 
er than their own. 

Discussion 

The NGA supports one of the highest densities 
of breeding raptors in the world (Marti et al. 
1993), and many previous studies have examined 
food habits of nesting raptors there (e.g., Marks 
and Marks 1981, Marks and Doremus 1988, Marti 
1988, Steenhof and Kochert 1988). However, die- 
tary habits and trophic relationships of Burrowing 
Owls remain the least well-understood of raptors 
breeding in the NCA (Marti pers. comm.). Thus, 
our study filled an important knowledge gap in 


raptor ecology within the NCA. Our study found: 
(1) no one species dominated the vertebrate com- 
ponent of Burrowing Owl diets, unlike owls in oth- 
er regions; (2) diets differed by habitat, most no- 
tably that montane voles and crickets were 
important prey for agricultural nests, but they were 
not part of the diet for nonagricultural nests; and 
(3) the food-niche breadth of Burrowing Owls in 
the NCA is broader than previously estimated. 

Burrowing Owl Diet in the NCA. Burrowing 
Owls are considered opportunistic predators 
(Gleason and Craig 1979, Green et al. 1993, Haug 
et al. 1993), and the wide variety of prey owls con- 
sumed in our study area is consistent with this no- 
tion. Similar to studies in Colorado (Marti 1974), 
Saskatchewan (Haug 1985), and the Idaho Nation- 
al Engineering Laboratory (INEEL) in Idaho 
(Gleason and Craig 1979), invertebrates represent- 
ed approximately 90-95% of prey items in regur- 
gitated pellets, but they constituted only 20-30% 
of biomass of prey. In contrast, Olenick (1990), in 


434 


Moulton et al. 


VoL. 39, No. 4 


Table 3. Mean (±SE) percent number and percent biomass per nest of vertebrate (taxa) and invertebrate (class/ 
order) prey detected in pellets of Burrowing Owls nesting in agricultural {N = 34) and nonagricultural {N = 19) 
habitats of southwestern Idaho, 2001-02. 


Habitat 


Prey Category Agricultural Nonagricultural Z® P-value 


Percent Number 


Mammal 

4.9 


0.8 

10.1 

-h 

1.1 

3.01 

0.003* 

Bird 

0.1 


0.1 

0.3 

-h 

0.1 

0.22 

0.823 

Reptile and Amphibian 

0.1 

-h 

0.1 

0.1 


0.1 

-0.38 

0.701 

Arachnid 

6.7 

± 

1.7 

26.5 


2.3 

4.98 

<0.001* 

Orthopteran 

40.0 


3.5 

16.7 


4.7 

-3.61 

<0.001* 

Goleopteran 

47.6 

± 

3.2 

46.1 


4.3 

-0.29 

0.774 

Total vertebrates 

5.0 


0.8 

10.3 

+ 

1.1 

3.05 

0.002* 

Total invertebrates 

95.0 

+ 

0.8 

89.7 


1.1 

-3.05 

0.002* 

Percent Biomass 

Mammal 

70.7 

+ 

3.1 

76.9 

+ 

4.1 

1.03 

0.303 

Bird 

2.3 

± 

1.0 

2.0 


1.3 

0.02 

0.988 

Reptile and Amphibian 

2.8 


1.5 

0.3 


2.0 

-0.51 

0.613 

Arachnid 

3.7 


1.1 

10.9 


1.4 

4.12 

<0.001* 

Orthopteran 

12.7 


1.9 

4.0 

-1- 

2.6 

-3.24 

0.001* 

Goleopteran 

7.5 

± 

0.8 

5.7 


0.1 

-0.99 

0.321 

Total vertebrates 

75.8 

-1- 

2.6 

79.1 

-K 

3.5 

0.96 

0.340 

Total invertebrates 

24.2 

± 

2.6 

20.9 


3.5 

-0.96 

0.340 


® Data were compared using Wilcoxon’s ranked sums tests. 

* Significant based on sequential Bonferroni corrections adjusted from an original alpha level of 0.05 for a total of 16 comparisons 


Table 4. Mean (±SE) percent number and percent biomass per nest of rodent species detected in pellets of Bur- 
rowing Owls nesting in agricultural {N = 34) and nonagricultural {N = 19) habitats of southwestern Idaho, 2001-02. 


Habitat 


Prey Species Agricultural Nonagricultural P-value 


Percent Number 


Spermophilus mollis 

0.1 


0.1 

0.5 


0.2 

1.83 

0.067 

Thomomys townsendii 

0.4 


0.1 

0.0 

-h 

0.1 

-2.72 

0.007 

Perognathus parvus 

0.7 

-+- 

0.6 

5.6 


0.8 

4.23 

<0.001* 

Dipodomys ordii 

0.4 

-+- 

0.2 

1.2 

-h 

0.3 

1.67 

0.095 

Peromyscus maniculatus 

1.4 

± 

0.5 

2.1 

-1- 

0.7 

-1.43 

0.153 

Mus musculus 

0.2 


0.1 

0.0 


0.1 

-2.25 

0.025 

Microtus montanus 

1.4 


0.3 

0.0 


0.4 

-4.32 

<0.001* 

Percent Biomass 

Spermophilus mollis 

5.3 

± 

4.1 

0.2 

-h 

0.1 

1.86 

0.063 

Thomomys townsendii 

1.9 

H- 

4.1 

0.0 

-1- 

5.4 

-2.72 

0.007 

Perognathus parvus 

4.9 


2.6 

26.0 


3.4 

4.00 

<0.001* 

Dipodomys ordii 

4.9 


2.8 

17.5 


3.8 

1.79 

0.073 

Peromyscus maniculatus 

10.5 


3.1 

12.3 


4.2 

-1.05 

0.294 

Mus musculus 

1.9 


0.6 

0.0 

± 

0.8 

-2.25 

0.025 

Microtus montanus 

20.6 


3.1 

0.0 


4.2 

-4.32 

<0.001* 


Data were compared using Wilcoxon’s ranked sums tests. 

* Significant based on sequential Bonferroni corrections adjusted from an original alpha level of 0,05 for a total of 14 comparisons 


December 2005 


Burrowing Owl Food Habits in Idaho 


435 


Table 5. Mean (±SE) percent number and percent biomass per nest of Arachnid orders and Orthopteran families 
detected in pellets of Burrowing Owls nesting in agricultural {N = 34) and nonagricultural {N = 19) habitats of 
southwestern Idaho, 2001-02. 

Prey Order/Family 

Habitat 

Agricultural 

Non agricultural 


P-VALUE 

Percent Number 
Arachnida 
Scorpiones 

3.3 ± 1.1 

10.3 ± 1.5 

2.22 

0.026 

Solpugida 

3.4 ± 1.6 

16.2 ± 2.1 

4.04 

<0.001* 

Orthoptera 

Acrididae 

1.8 ± 0.8 

4.9 ± 1.0 

2.81 

0.005* 

Gryllidae 

34.8 ± 3.5 

2.3 ± 4.7 

-5.43 

<0.001* 

Percent Biomass 

Arachnida 

Scorpiones 

2.0 ± 1.0 

6.0 ± 1.4 

1.72 

0.086 

Solpugida 

1.7 ± 0.6 

4.9 ± 0.8 

3.68 

<0.001* 

Orthoptera 

Acrididae 

0.5 ± 0.2 

1.1 ± 0.2 

2.38 

0.017* 

Gryllidae 

11.4 ± 1.9 

1.0 ± 2.5 

-5.26 

<0.001* 


Data were compared using Wilcoxon’s ranked sums tests. 

* Significant based on sequential Bonferroni corrections adjusted from an original alpha level of 0.05 for a total of four comparisons 
each for Arachnida and Orthoptera. 


southeastern Idaho, reported that invertebrates 
represent only 60% of the number of prey items 
and less than 3% of the biomass, and owls in the 
Imperial Valley, California, feed almost exclusively 
on invertebrates (York et al. 2002). 

Although invertebrates generally constitute a 
large percentage of prey Burrowing Owls consume, 
the orders and families that are most common in 
the diet vary among regions. For example, Cole- 
opterans were the most abundant invertebrate spe- 
cies in our study, as well as in Colorado (Marti 
1974), Washington (Green et al. 1993), and 
Oregon (Green et al. 1993), whereas Jerusalem 
crickets (Stenopelmatus spp.) were the most impor- 
tant invertebrate prey species, in terms of biomass, 
for Burrowing Owls in Oregon (Green et al. 1993), 
California (Thomsen 1971), and southeastern Ida- 
ho (Gleason and Craig 1979). 

Vertebrates accounted for most of the biomass 
in our study, but no one vertebrate species domi- 
nated the diet. Percent biomass of montane voles 
(17%), pocket mice (16%), pocket gophers (16%), 
and deer mice (14%) were similar. In contrast, Mi- 
crotus sp. were the predominant vertebrate prey 
item in Montana (Holt et al. 2001) and represent- 
ed 80% of biomass in owl diets in southeastern Ida- 


ho (Olenick 1990), and pocket mice dominated 
rodent prey in Oregon (97%; Green 1983). This 
lack of a dominant vertebrate prey may indicate a 
diverse prey base in our study area (Moulton et al. 
in press). 

Agricultural versus Nonagricultural Nests. Com- 
parisons of pellet remains from Burrowing Owl 
nests in agricultural and nonagricultural areas re- 
vealed different prey composition, species rich- 
ness, species evenness, and food-niche breadth. Al- 
though both habitats had similar biomass of 
vertebrates, nonagricultural areas had greater 
numbers of rodent prey. In contrast, owls nesting 
adjacent to agricultural fields in southeastern Ida- 
ho had a higher proportion of rodents in their diet 
than those nesting in more natural areas (Gleason 
1978). Agricultural nests had a higher proportion 
of invertebrates than nonagricultural nests, which 
resulted from the high numbers of crickets present 
in pellets from agricultural nests. Crickets were 
rare in pellets of owls nesting in nonagricultural 
habitats. Moulton et al. (in press) reported greater 
prey consumption by Burrowing Owls nesting near 
agricultural fields in the NCA; this difference pri- 
marily resulted from greater invertebrate prey in 
agricultural habitats. While some have suggested 


436 


Moulton et ai.. 


VoL. 39, No. 4 


Table 6. Percent number, percent biomass, and total number of cached and other uneaten prey remains docu- 
mented at 43 Burrowing Owl nests in southwestern Idaho, 2001-02. 


Prey Category 

Percent No. 

Percent Biomass 

Total No. 

Mammals 

73.70 

87.67 

297 

^Ivilagus nuttallii 

0.25 

1.03 

1 

Thomomys townsendii 

10.91 

44.73 

44 

Dipodomys ordii 

11.41 

13.01 

46 

Perognathus parvus 

2.48 

0.77 

10 

Mus musculus 

2.98 

1.17 

12 

Mouse species® 

18.86 

5.86 

76 

Microtus montanus 

26.80 

21.10 

108 

Birds 

4.47 

8.09 

18 

Eremophila alpestris 

0.25 

0.09 

1 

Sturnus vulgaris 

0.74 

1.22 

3 

Sturnella neglecta 

0.50 

0.41 

2 

Passerine sp,^ 

0.25 

0.15 

1 

Athene cuniculana — juv. 

2.23 

4.16 

9 

A. cunicularia — adult 

0.25 

1.03 

1 

Raptor sp.'^ 

0.25 

1.03 

1 

Amphibians 

8.68 

3.60 

35 

Bufo woodhousei 

8.68 

3.60 

35 

Reptiles 

0.74 

0.29 

3 

Pituophis catenifer 

0.74 

0.29 

3 

Scolopendromorpha 

0.50 

0.00 

2 

Arachnids 

10.92 

0.34 

44 

Scorpiones 

10.67 

0.33 

43 

Solpugida 

0.25 

0.01 

1 

Orthopterans 

0.50 

0.01 

2 

Acrididae 

0.25 

0.00 

1 

Gryllidae 

0.25 

0.00 

1 

Total vertebrates 

87.59 

99.65 

353 

Total invertebrates 

12.41 

0.35 

50 

Total 



403 


Likely P. parvus, R. megalotis, P. maniculatus, or M. musculus. 
Likely Eremophila alpestris or Sturnella neglecta. 

Small juvenile hawk or Prairie Falcon {Falco mexicanus) . 


that Burrowing Owls associate with irrigated agri- 
culture because of the high abundance of montane 
voles (Gleason 1978, Rich 1986), presence of high 
numbers of invertebrate prey in the diet of owls in 
agricultural habitat may indicate an overlooked im- 
portance of invertebrate prey to breeding Burrow- 
ing Owls in these areas. 

Agricultural nests also had greater species rich- 
ness than nonagricultural nests. Common rodent 
species in agricultural habitats, such as montane 
voles, were not in pellets of nonagricultural nests 
and likely were not available in that habitat type. 
However, nonagricultural nests had greater species 


evenness than agricultural nests. This greater spe- 
cies evenness likely contributed to our finding that 
diets of owls nesting in nonagricultural areas had 
a broader food-niche (i.e., greater diversity), as 
Simpson’s diversity index can be greatly influenced 
by species evenness. 

Narrower food-niche breadths of Burrowing 
Owls nesting near agricultural fields may indicate 
a more specialized diet. As MacArthur and Pianka 
(1966) proposed, one expects a species to special- 
ize when prey availability is high (i.e., a productive 
environment), and thus search time is low. A spe- 
cies will generalize in unproductive environments 


December 2005 


Burrowing Owl Food Habits in Idaho 


437 


where search times are high. Therefore, if owls in 
agricultural areas exhibit more specialized diets 
relative to owls in nonagricultural areas, we pro- 
pose that owls nesting in agricultural areas are ex- 
periencing greater prey availability. This is consis- 
tent with suggestions by previous researchers 
(Gleason 1978, Rich 1986, Moulton et al. in press) 
that Burrowing Owls associate with agriculture be- 
cause of increased prey. However, further research 
is needed to determine if the narrower food-niche 
breadth of owls in agricultural areas results from 
greater prey availability, where owls can be selec- 
tive, or lower prey diversity. 

Food-niche Breadth of Burrowing Owls in the 
NCA. Prior to our study. Burrowing Owls in the 
NCA were thought to have a very narrow food- 
niche breadth compared to other raptor species 
breeding there. Marti et al. (1993) estimated food- 
niche breadth of Burrowing Owls to be only 2.43, 
which was the narrowest food-niche breadth of all 
12 raptor species studied. In contrast, food-niche 
breadth of Burrowing Owls in our study was 4.22 
± 0.22, which ranks Burrowing Owls seventh in 
terms of food-niche breadth (first being the broad- 
est). This disparity may be explained in part by 
smaller sample sizes in Marti et al. (1993) com- 
bined with different levels of identification; that is, 
the 1993 study identified invertebrate prey to or- 
der, whereas we identified invertebrates to family 
when possible. Because this difference in prey level 
identification would only affect the food-niche 
breadth estimates of a species whose diet has a 
large invertebrate component, only Burrowing Owl 
estimates likely would be affected. 

Compared to other raptors breeding within the 
NCA, our study estimated food-niche breadths of 
Burrowing Owls to be similar to Golden Eagles 
{Aquila chrysaetos; 4.07) and Long-eared Owls {Asio 
otus; 4.79; Marti et al. 1993). However, Burrowing 
Owl diet composition is more similar to American 
Kestrels {Falco sparverius), which also frequently 
prey on invertebrates (Marti et al. 1993). In fact, 
Burrowing Owls and American Kestrels are the 
only two raptor species in the NCA for which in- 
vertebrate prey comprises >1% of the diet (in 
terms of biomass: 23% and 5%, respectively). 

Acknowitdgments 

We thank T. Booms, K. Donohue, C. Eaton, J. Egbert, 
G. Fairhurst, K. Hasselblad, S. Lysne, R. Medeck, J. 
Rausch, J. Soules, M. Troxler, and J. Urbanic for assis- 
tance with fieldwork. Financial and logistical support was 
provided through an ERA STAR Fellowship to C. Moul- 
ton, grants from the Bureau of Land Management and 


Idaho Department of Fish and Game, and the Depart- 
ment of Biology and Raptor Research Center at Boise 
State University. J. Clark, J. Doremus, and J. Sullivan fa- 
cilitated our work in the Snake River Birds of Prey Na- 
tional Conservation Area, and M. Fuller, Director of the 
Raptor Research Center, was helpful in numerous ways 
Finally, we thank J. Bednarz, C.S. Houston, J. Munger, T. 
Rich, and I. Robertson for helpful comments on previous 
versions of our manuscript. 

Literature Cited 

Alatalo, R.V. 1981. Problems in the measurement of 
evenness in ecology. Oikos 37:199-204. 

Bellocq, M.A. 1997. Ecology of the Burrowing Owl in 
agrosystems of central Argentina. Pages 52—57 m J.L. 
Lincer [Ed.], The Burrowing Owl, its biology and 
management including the proceedings of the first 
international Burrowing Owl symposium. Raptor Re- 
search Report 9. 

Belthoff, J.R. AND B.W. Smith. 2003. Patterns of artificial 
burrow occupancy and reuse by Burrowing Owls in 
Idaho. Wildl. Soc. Bull. 31:138-144. 

Brady, R.S. 2004. Nest-lining behavior, nest microclimate, 
and nest defense behavior of Burrowing Owls. M.S. 
thesis, Boise State University, Boise, ID U.S.A. 
Carlson, C.A. 1985. Wildlife and agriculture; can they 
coexist? J. Soil Water Conserv. 40:263-266. 

Clayton, K.M. and J.K. Schmutz. 1999. Is the decline of 
Burrowing Owls (Speotyto cunicularia) in prairie Can- 
ada linked to changes in Great Plains ecosystems? Bird 
Conserv. Int. 9:163-185. 

DeSante, D.F., E.D. Ruhlen, and D.K. Rosenberg. 2004 
Density and abundance of Burrowing Owls in the ag- 
ricultural matrix of the Imperial Valley, California 
Stud. Avian Biol. 27:116-119. 

Gervais, J.A., D.K. Rosenberg, D.M. Fry, L. Trulio, and 
K.K. Sturm. 2000. Burrowing Owls and agricultural 
pesticides: evaluation of residues and risks for three 
populations in California, U.S.A. Environ. Toxicol 
Chem. 19:337-343. 

Gleason, R.L. 1978. Aspects of the breeding biology of 
Burrowing Owls in southeastern Idaho. M.S. thesis. 
University of Idaho, Moscow, ID U.S.A. 

AND T.H. Craig. 1979. Food habits of Burrowing 

Owls in southeastern Idaho. Great Basin Nat. 39:273— 
276. 

Green, G.A. 1983. Ecology of breeding Burrowing Owls 
in the Columbia Basin, Oregon. M.S. thesis, Oregon 
State University, Corvallis, OR U.S.A. 

, R.E. Fitzner, R.G. Anthony, and L.E. Rogers. 

1993. Comparative diets of Burrowing Owls in 
Oregon and Washington. Northwest Sci. 67:88-93. 
Haug, E.A. 1985. Observations on the breeding ecology 
of Burrowing Owls in Saskatchewan. M.S. thesis. Uni- 
versity of Saskatchewan, Saskatoon, Canada. 

, B.A. Millsap, and M.S. Martell. 1993. Burrow- 
ing Owl {Speotyto cunicularia). In A. Poole and F. Gill 
[Eds.], The birds of North America, No. 61. The 
Academy of Natural Sciences, Philadelphia, PA and 


438 


Moulton et al. 


VoL. 39, No. 4 


The American Ornithologists’ Union, Washington, 
DC U.S.A. 

AND L.W. Oliphant. 1990. Movements, activity 

patterns, and habitat use of Burrowing Owls in Sas- 
katchewan./. Wildl. Manage. 54:27-35. 

Hill, M.O. 1973. Diversity and evenness: a unifying no- 
tation and its consequences. Ecology 54:427-432. 

Hironaka, M., M.A. Fosberg, and A.H. Winward. 1983. 
Sagebrush-grass habitat types of southern Idaho. Bul- 
letin Number 35. Forest, Wildlife and Range Experi- 
ment Station, University of Idaho, Moscow, ID U.S.A. 

Holt, D.W., W.D. Norton, and W.C. Atkinson 2001. 
Breeding season food habits of Burrowing Owls in 
south-central Montana. Intermt. J. Sci. 7:63—69. 

Jahn, L.R. and E.W. Schenck. 1991. What sustainable ag- 
riculture means for fish and wildlife. / Soil Water Con- 
serv. 46:251-254. 

James, P.C. and R,M. Espie. 1997. Current status of the 
Burrowing Owl in North America: an agency survey. 
Pages 3-5 mJ.L. Lincer [Ed.]. The Burrowing Owl, 
its Biology and Management including the proceed- 
ings of the first international Burrowing Owl sympo- 
sium. Raptor Research Report 9. 

and G.A. Fox. 1987. Effects of some insecticides 

on productivity of Burrowing Owls. Blue Jay 45:65-71. 

Klute, D.S., L.W. Ayers, M.T. Green, W.H. Howe, S.L. 
Jones, J.A. Shaffer' S.R. Sheffield, and T.S. Zimmer- 
man. 2003. A status assessment and conservation plan 
for the western Burrowing Owl in the United States. 
U.S. Department of Interior, Fish and Wildlife Ser- 
vice, Biological Technical Publication FWS/BTP- 
R6001-2003, Washington, DC U.S.A. 

Leptich, D. 1994. Agricultural development and its influ- 
ence on raptors in southern Idaho. Northwest Sci. 68: 

167-171. 

MacArthur, R.H. and E.R. Pianka. 1966. On optimal use 
of a patchy environment. Am. Nat. 100:603-609. 

Marks, J.S. and J.H. Doremus. 1988. Breeding-season 
diet of Northern Saw-whet Owls in southwestern Ida- 
ho. Wilson Bull. 100:690-694. 

AND V.A. Marks. 1981. Comparative food habits 

of the Screech Owl and Long-eared Owl in south- 
western Idaho. Murrelet 62:80-82. 

Marti, C.D. 1974. Feeding ecology of four sympatric 
owl s . Condor 7 6 : 45—6 1 . 

. 1987. Raptor food habits studies. Pages 67-80 in 

B.A. Pendleton, B.A. Millsap, K.W. Kline, and D.M. 
Bird [Eds.], Raptor management techniques manual. 
National Wildlife Federation Scientific and Technical 
Series No. 10, Washington, DC U.S.A. 

. 1988. A long-term study of food-niche dynamics 

in the Common Barn Owl: comparisons within and 
between populations. Can. J. Zool. 66:1803—1812. 

, K. Steenhof, M.N. Kochert, and J.S. Marks. 

1993. Community trophic structure: the roles of diet. 


body size, and activity time in vertebrate predators. 
Oikos 67:6—18. 

Moulton, C.E. 2003. Ecology of Burrowing Owls in 
southwestern Idaho: association with agriculture, food 
habits, and territorial behavior. M.S. thesis, Boise State 
University, Boise, ID U.S. A. 

, R.S. Brady, and J.R. Belthoff. In press. Associ- 
ation between wildlife and agriculture: underlying 
mechanisms and implications in Burrowing Owls. / 
Wildl. Manage. 

Murphy, M.T. 2003. Avian population trends within the 
evolving agricultural landscape of eastern and central 
United States. Auk 120:20-34. 

Olenick, B.E. 1990. Breeding biology of Burrowing Owls 
using artificial nest burrows in southeastern Idaho. 
M.S. thesis, Idaho State University, Pocatello, ID 
U.S.A. 

Peterjohn, B.G. 2003. Agricultural landscapes: can they 
support healthy bird populations as well as farm prod- 
ucts? Auk 120:14-19. 

Rice, W.R. 1989. Analyzing tables of statistical tests. Evo- 
lution 43:223-225. 

Rich, T. 1986. Habitat and nest-site selection of Burrow- 
ing Owls in the sagebrush steppe of Idaho. /. Wildl. 
Manage. 50:548-555. 

Rosenberg, D.K and K.L. Hai.ey. 2004. Ecology of Bur- 
rowing Owls in the agroecosystem of the Imperial Val- 
ley, California. Stud. Avian Biol. 27:120-135. 

Simpson, E.H. 1949. Measurement of diversity. Nature 
163:688. 

Smith, B.W. and J.R. Belthoff. 2001. Effects of nest di- 
mensions on use of artificial burrow systems by Bur- 
rowing Owls./. Wildl. Manage. 65:318-326. 

Smith, D.G. and J.R. Murphy. 1973. Breeding ecology of 
raptors in the eastern Great Basin of Utah. Brigham 
Young Univ. Sci. Bull. Biol. Ser. 18:69-76. 

Steenhof, K. 1983. Prey weights for computing percent 
biomass in raptor diets. Raptor Res. 17:15—27. 

AND M.N. Kochert. 1988. Dietary responses of 

three raptor species to changing prey densities in a 
natural environment./ Anim. Ecol. 57:37-48. 

Thomsen, L. 1971. Behavior and ecology of Burrowing 
Owls on the Oakland Municipal Airport. Condor 73: 
177-192. 

U.S. Department of the Interior. 1996. Effects of mil- 
itary training and fire in the Snake River Birds of Prey 
National Conservation Area. BLM/IDARNG Research 
Project Final Report. U.S. Geological Survey, Biol 
Res. Div., Snake River Field Station, Boise, ID U.S.A. 

York, M.M., D.K. Rosenberg, and K.K. Sturm. 2002. Diet 
and food-niche breadth of Burrowing Owls {Athene cu- 
nicularia) in the Imperial Valley, California. West. N 
Am. Nat. 62:280-287. 

Zar, J.H. 1999. Biostatistical analysis, 4th Ed. Prentice 
Hall, Upper Saddle River, NJ U.S.A. 

Received 1 December 2003; accepted 3 July 2005 


J Raptor Res. 39(4) :439-444 
© 2005 The Raptor Research Foundation, Inc. 


RED-TAILED HAWK DIETARY OVERLAP WITH NORTHERN 
GOSHAWKS ON THE KAIBAB PLATEAU, ARIZONA 

Angela E. Gatto^ 

Bureau of Land Management, Lake Havasu, 2610 Sweetwater Avenue, Lake Havasu City, AZ 86406 U.S.A. 

Teryl G. Grubb 

Rocky Mountain Research Station, 2500 South Pine Knoll Drive, Flagstaff, AZ 86001 U.S.A. 

Carol L. Chambers 

School of Forestry, College of Ecosystem Science and Management, Northern Arizona University, P.O. Box 15018, 

Flagstaff, AZ 86001 U.S.A. 

Abstract. — ^We determined food habits of Red-tailed Hawks {Buteo jamaicensis) for comparison with 
published information for Northern Goshawks {Accipiter gentilis) to evaluate potential competition on 
the Kaibab Plateau, Arizona. We collected prey remains and pellets from 42 Red-tailed Hawk nests at 
the end of the nesting season between August-October 1998-2001, and opportunistically from below 
nest trees during site visits, May-July 2000-01. We identified 478 prey items, including 17 mammal, 7 
bird, and 2 reptile species. Prey species frequency did not vary among years (P = 0.3), across habitat 
types (P = 0.8), or by collection technique (P = 0.4). Annual food niche breadth for Red-tailed Hawks 
averaged 0.57. Published mean niche breadth for Northern Goshawks was 0.32, supporting that Red- 
tailed Hawks were feeding generalists, while Northern Goshawks were more specialized. However, 48% 
of Red-tailed Hawk diet on the Kaibab Plateau consisted of species comprising a major portion of the 
documented diet of Northern Goshawks, including Nuttall’s cottontail {Sylvilagus nuttallii), golden-man- 
tled ground squirrel {Spermophilus lateralis lateralis), rock squirrel (5. variegates grammurus) , and Northern 
Elicker ( Colaptes auratus) . Because raptor communities with high dietary overlap and lack of prey par- 
titioning show food-limited nesting success, greater agnonistic behavior, and territoriality, Red-tailed 
Hawks could be negatively affecting Northern Goshawks on the Kaibab Plateau. 

Key Words: Red-tailed Hawk, Buteo jamaicensis; Northern Goshawk, Accipiter gentilis; competition', diet, 

food habits; food niche breadth; foraging. 


SOBRELAPAMIENTO DE LA DIETA DE BUTEO fAMAICENSlSY ACCIPITER GPAP/L/SEN LAMESETA 
DE KAIBAB, ARIZONA 

Resumen. — Determinamos la dieta de Buteo jamaicensis en la planicie de Kaibab, Arizona y la compara- 
mos con informacion publicada sobre Accipiter gentilis con la finalidad de evaluar la posible existencia 
de competencia entre estas dos especies. Recolectamos restos de presas y egagropilas de 42 nidos de B. 
jamaicensis al final de la temporada reproductiva entre agosto y octubre de 1998-2001, y de forma 
oportunista debajo de arboles de anidacion durante visitas a la zona de estudio realizadas entre mayo 
y julio de 2000 y 2001. Identificamos 478 tipos de presas, induyendo 17 especies de mamiferos, 7 de 
aves y 2 de reptiles. La frecuencia de las especies de presa no vario entre anos (P = 0.3), tipos de 
habitat (P = 0.8) y tecnicas de colecta (P = 0.4). La amplitud anual del nicho alimentario de B. 
jamaicensis promedio 0.57. Registros publicados indican que la amplitud del nicho alimentario de A. 
gentilis es 0.32, lo cual sugiere que B. jamaicensis es una especie generalista, mientras que A. gentilis es 
mas especializada. Sin embargo, el 48% de la dieta de B. jamaicensis en la planicie de Kaibab consiste 
de especies que forman la mayor parte de la dieta documentada para A. gentilis, entre las que se 
encuentran Sylvilagus nuttallii, Spermophilus lateralis lateralis, S. variegates grammurus y Colaptes auratus. 
Debido a que las comunidades de rapaces que presentan un alto grado de superposicion en su dieta y 
en las que no hay reparticion de presas muestran un bajo exito de anidacion debido a falta de alimento, 


^ Email address: angela_gatto@blm.gov 


439 


440 


Gatto et ai.. 


VoL. 39, No. 4 


un mayor comportamiento antagonico y mayor territorialidad, B. jamaicensis podria estar afectando 
negativamente a A. gentilis en la planicie de Kaibab. 

[Traduccion del equipo editorial] 


Fire suppression, timber harvesting, and live- 
stock grazing over the past 100 yr have caused 
changes in southwestern forests, including those 
on the Kaibab Plateau in northwestern Arizona 
(e.g.. Weaver 1951, Cooper 1960, Covington and 
Moore 1994). Decreasing the quality and quantity 
of climax forest can favor habitat generalists (Carey 
1984) such as Red-tailed Hawks {Buteo jamaicensis) , 
which may out compete species dependent on late 
successional forests, such as Northern Goshawks 
{Accipiter gentilis; Andersen et al. 2003). Creation of 
259-ha buffer zones on the plateau by the Kaibab 
National Forest, although maintaining the quality 
of old-growth for nesting sites, does not ensure suf- 
ficient quantity or quality of old-growth forest be- 
yond these protected areas (Reynolds et al. 1992). 
Red-tailed Hawks often nest in abandoned North- 
ern Goshawk nests after timber harvesting (Crock- 
er-Bedford 1990). Beyond old-growth-managed 
buffer zones, open areas may favor the hunting 
style, and thus, increase foraging habitat of Red- 
tailed Hawks more than Northern Goshawks. Ex- 
ploring the potential relationship between these 
two species facilitates a better understanding of 
how upper level avian predators have adjusted to 
human-caused alterations in southwestern forest 
ecosystems. Further, this understanding could aid 
resource managers in maintaining a viable North- 
ern Goshawk population on the Kaibab Plateau. 

Competition occurs when use or defense of a 
resource by one individual reduces the availability 
of that resource to other individuals (Gill 1990). 
Interspecific competition occurs when individuals 
of coexisting species require a resource that is in 
limited supply relative to their needs such that sur- 
vival or reproduction of at least one species is de- 
creased (Ricklefs and Schluter 1993). On the Kai- 
bab Plateau, Red-tailed Hawks and Northern 
Goshawks, based on their proximity and similar 
habitat use, could be competing for nest sites, for- 
aging areas, or prey, despite potential partitioning 
resulting from morphological and behavioral dif- 
ferences between species (Ballam 1984, Squires 
and Reynolds 1997). In this study, we focused on 
determining the diet of Red-tailed Hawks for com- 
parison with published information on Northern 
Goshawks to evaluate potential overlap. We hypoth- 


esized that Red-tailed Hawks and Northern Gos- 
hawks, because of their close proximity and use of 
similar nesting and foraging habitat, could also be 
using the same prey species, creating the potential 
for competition. 

Study Area 

The Kaibab Plateau, on the North Rim of the Grand 
Canyon, is located in northwestern Arizona within the 
Kaibab National Forest. The plateau encompasses 2980 
km^ above 1830 m elevation. Our study area was confined 
to 1732 km^ above the 2075 m contour to be consistent 
with concurrent long-term Northern Goshawk research 
on the plateau (Reynolds et al. 1994). Goshawk and Red- 
tailed Hawk nest sites are interspersed throughout this 
study area. Vegetation on the plateau consists of ponde- 
rosa pine {Finns ponderosa) forest between 2075-2500 m; 
mixed-conifer forest (ponderosa pine, Douglas-fir [Pseu- 
dotsuga menziesii], white fir [Abies concolor], blue spruce 
[Picea pungens] , and quaking aspen [Populus tremuloides ^ ) 
between 2500-2650 m; and Engelmann spruce-subalpine 
fir {Picea engelmannii-Abies lasiocarpa) forest between 
2650-2800 m (Rasmussen 1941, White and Vankat 1993) 

Methods 

Prey Remains and Pellet Identification. We collected 
prey remains and regurgitated pellets from all known oc- 
cupied Red-tailed Hawk nests {N = 21, 24, 32, 11 for 
1998-2001, respectively), from below and in the nest, at 
the end of the nesting season, August-October 1998- 
2001, and opportunistically during the nesting season, 
May-Jnly 2000-01. Collections from each nest were sep- 
arated by >3 d, and end-of-season samples were collected 
>30 d after previous collections. We treated each collec- 
tion of pellets and prey remains from an occupied nest 
as one sample for that nest site. We assumed prey species 
identified in each sample had been consumed since the 
last sample and represented new prey. 

To identify prey, we separated samples into bones/frag- 
ments, feathers, and hair. Bones and feathers were iden- 
tified to the lowest taxon possible, through comparison 
with S. Bayard’s reference prey collection stored at the 
Rocky Mountain Research Station, Fort Collins, CO. We 
identified hairs using keys (Williams 1938, Stains 1958, 
Moore et al. 1974) and by comparing hairs directly with 
samples from the Northern Arizona University, Depart- 
ment of Biological Sciences’ collection and the private 
collection of H.E. Graham (Flagstaff, AZ U.S.A.). The 
characteristics we considered included color banding, 
shape, presence of a hair shield, and configuration of a 
medulla if present (Moore et al. 1974). 

Determining diet via indirect means requires cautious 
interpretation because of inherent biases associated with 
each method (Lewis et al. 2004); however, identification 
of a prey species in pellets and prey remains is an abso- 
lute indication of presence. Therefore, we recorded one 
occurrence whenever a species was found in a sample. 


December 2005 


Red-tailed Hawk Diet 


441 


We then pooled samples across years for each nest and 
summed the number of times a prey species occurred. 
This provided a conservative estimate of the relative im- 
portance of prey species consumed by Red-tailed Hawks 
at each nest. We also calculated the total number and 
percent of nests at which each prey species occurred. 

Prey Species Dissimilarity. Dyer (1978) developed a lin- 
ear statistical analysis for comparing species dissimilarity 
that can be used in conjunction with any species dissim- 
ilarity index and is designed for data sets that involve 
both multiple species and multiple environmental vari- 
ables. Total species dissimilarity is divided into compo- 
nents with one component being assigned to each envi- 
ronmental variable or interaction of environmental 
variables: 

= Po + + W' + - + 

where D^yis the (dis) similarity between observations zand 
j, 5,yh) is a known function of i and j which corresponds 
to the environmental variable or interaction, Pj is an 
unknown parameter which represents the contribution 
of the environmental variable or interaction to the 
total dissimilarity, and cij is an error term with an ex- 
pected value of 0. 

We used this linear model with the Jaccard dissimilarity 
index (Krebs 1998) to estimate whether the species iden- 
tified in prey remains and pellets of Red-tailed Hawks 
varied among years (1998-2001), vegetation types (pon- 
derosa pine only, mixed conifer with pine dominant, or 
mixed conifer only), or collection techniques (end of 
season samples from nest or opportunistic samples col- 
lected from below nest trees during nesting season). Jac- 
card’s index is specifically designed for presence-absence 
data, and because, by definition, rare species are typically 
absent, rare species have little influence on the value of 
the index (Krebs 1998). The Jaccard dissimilarity index 
(D^J) is calculated by: 

Djy = a/ {a + b + c) 

where a is the number of binary characteristics present 
m sample i and sample j, b is the number present only 
in i, and c is the number present only in j. We estimated 
Dyer’s model using the Jaccard index. We used permu- 
tation methods (Edington 1995) to estimate a signifi- 
cance level (P-value) for each environmental variable. 
Statistical tests were significant if P < 0.05. 

Niche Breadth. Niche breadth and niche overlap are 
widely applied to analysis of foraging and community 
ecology to estimate competition (Greene and Jaksic 
1983). Niche breadths were calculated according to Lev- 
ins’ (1968) equation: 

P = for i = \ to n 

Zj Pi^ 

where pi is the proportion of Red-tailed Hawk nests with 
the taxon present. The value of p varies from 1 to n, 
where n is the number of taxa. If prey taxa occur equally 
among all nests, then p = ra. Niche-breadth values were 
standardized and converted to a fraction ranging from 0 
to 1 by the equation: 

Pstandard = (P “ ^) / {n ~ 1). 

To calculate niche breadth, we created a prey species 


list based on pooled prey species present across all oc- 
cupied Red-tailed Hawk nests by year. This yielded a sep- 
arate food niche breadth measure for each of the four 
study years. We also calculated a 95% confidence interval 
for Red-tailed Hawk niche breadth on the Kaibab Plateau 
for comparison with calculated niche breadths for North- 
ern Goshawks from other studies. 

Resuits 

Pellet and Prey Remain Analysis. Considering 
each visit’s collection of pellets and prey remains 
as one discreet sample, we obtained 140 prey sam- 
ples {N — 80 end of nesting season, 1998-2001; N 
= 60 opportunistic during nesting season, 2000- 

01) , consisting of 1-10 collections from 42 nests {x 
= 3.3). We identified 478 prey items, including at 
least 17 mammal, seven bird, and two reptile spe- 
cies (Table 1). The number of species at any one 
nest site ranged from 1-6. For all 4 yr combined, 
mammals represented 72% of Red-tailed Hawk 
diet and birds represented 27% (Table 1). Only 
two reptile species were identified; they did not 
contribute greatly to the overall diet (<1%). Six 
species accounted for 67% of prey by frequency of 
occurrence: Nuttall’s cottontail {Sylvilagus nuttallu, 
17.6%), Kaibab squirrel (Sciurus aberti kaibabensis, 
7.7%), rock squirrel {Spermophilus variegatus gram- 
murus, 10.0%), golden-mantled ground squirrel {S. 
lateralis lateralis, 10.3%), Northern Flicker {Colaptes 
auratus, 10.7%), and Steller’sJay {Cyanocitta stellen, 
10.3%; Table 1). Rare occurrences of porcupine 
(Erethizon dorsatum) , coyote ( Canis latrans) , and 
mule deer {Odocoileus hemionus) comprised <2% of 
overall Red-tailed Hawk diet. 

Dissimilarity Measures and Niche Breadth. Prey 
species frequency identified from pellet and prey 
remains did not vary among years {P = 0.3), across 
habitat types {P = 0.8), or by collection technique 
(P = 0.4). Dietary niche breadth for Red-tailed 
Hawks pooled across nests for each year was 0.58 
in 1998, 0.52 in 1999, 0.51 in 2000, and 0.65 in 
2001. The mean and 95% confidence interval for 
Red-tailed Hawks niche breadth for all four years 
combined was 0.57 ± 0.11, Calculated food niche 
breadths for Northern Goshawks for Arizona and 
several other western states did not fall within our 
confidence interval for Red-tailed Hawks (Table 

2) . Dietary overlap between the Red-tailed Hawk 
during this study and a Northern Goshawk diet 
from BoaPs (1993) earlier study in the same area 
was 55% (number of common species/sum of spe- 
cies recorded for both raptors). 


442 


Gatto et al. 


VoL. 39, No. 4 


Table 1. Frequency of prey species in 140 samples (478 identified prey items) of pellets and prey remains collected 
from 42 Red-tailed Hawk nests, May-October 1998-2001, on the Kaibab Plateau, Arizona, U.S.A. 


Prey Species 

Frequency 
(N = 478) 

Percent 

Frequency 

Nest 

Occurrence 
{N = 42) 

Percent 

Nest 

Occurrence 

Nuttall’s cottontail 
{Sylvilagus nuttallii) 

84 

17.6 

36 

85.7 

Golden-mantled ground squirrel 
{Spermophilus lateralis) 

49 

10.3 

26 

61.9 

Rock squirrel 

{Spermophilus variegatus grammurus) 

48 

10.0 

29 

69.0 

Kaibab squirrel 

{Sciurus aberti kaibabensis) 

37 

7.7 

21 

50.0 

Chipmunk 
{Eutamias sp.) 

27 

5.6 

20 

47.6 

Northern pocket gopher 
{Thomomys talpoides) 

22 

4.6 

21 

50.0 

Long-tailed vole 

{Microtus longicaudus) 

17 

3.6 

17 

40.5 

Red squirrel 

( Tamiasciurus hudsonicus) 

15 

3.1 

13 

31.0 

Mouse 

{Peromyscus sp.) 

13 

2.7 

9 

21.4 

Shrew 
{Sorex sp.) 

11 

2.3 

10 

23.8 

Black-tailed jackrabbit 
{Lepus californicus) 

8 

1.7 

8 

19.0 

Long-tailed weasel 

{Mustela frenata arizonenst) 

5 

1.0 

5 

11.9 

Mule deer 

(Odocoileus hemionus) 

5 

1.0 

5 

11.9 

Plains pocket mouse 
{Perognathus flavescens) 

3 

0.6 

2 

4.8 

Porcupine 

{Erethizon dorsatum) 

2 

0.4 

2 

4.8 

Ringtail 

{Bassariscus astutus) 

1 

0.2 

1 

2.4 

Coyote 

{Canis latrans) 

1 

0.2 

1 

2.4 

Northern Flicker 
{Colaptes auratus) 

51 

10.7 

28 

66.7 

Steller’s Jay 

( Cyanocitta stelleri) 

49 

10.3 

26 

61.9 

Clark’s Nutcracker 
{Nucifraga Columbiana) 

13 

2.7 

10 

23.8 

Unknown bird 

8 

1.7 

7 

16.7 

Common Raven 
{Corvus corax) 

3 

0.6 

3 

7.1 

Western Bluebird 
(Sialia mexicana) 

1 

0.2 

1 

2.4 

Hairy Woodpecker 
{Picoides villosus) 

1 

0.2 

1 

2.4 

Common Nighthawk 
( Chordeiles minor) 

1 

0.2 

1 

2.4 

Unknown snake 

3 

0.6 

2 

4.8 

Mountain short horned lizard^ 
{Phrynosoma douglassi) 

0 

0.0 

0 

0.0 


^ Mountain short horned lizard was positively identified as a prey species during observations, but was not found in the prey collec- 
tions. 


December 2005 


Red-tailed Hawk Diet 


443 


Table 2. Food niche breadth of nesting Northern Goshawks calculated from prey remains collected in Arizona and 
other western states. 


Location 

Number of 
Nests 

Food Niche 
Breadth 

Source 

Arizona 

20 

0.29 

Boal and Mannan (1994) 

California 

114 

0.41 

Bloom et al. (1986) 

New Mexico 

8 

0.36 

Kennedy (1991) 

Oregon 

4 

0.38 

Reynolds and Meslow (1984) 


Discussion 

Our results on the Kaibab Plateau were consis- 
tent with Red-tailed Hawks being feeding general- 
ists and preying primarily upon rabbits {Sylvilagus 
spp.), black-tailed jackrabbits {Lepus californicus) , 
and ground squirrels {Spermophilus spp; Preston 
and Beane 1993); however, Northern Flickers and 
Steller’s Jays were also frequent in this diet analysis. 
Previous observation and prey remains analyses for 
Northern Goshawks on the plateau indicated gos- 
hawks also preyed mostly upon rabbits and hares, 
tree and ground squirrels. Northern Flickers, and 
Steller’s Jays (Boal and Mannan 1994, Kaufmann 
et al. 1994, and Reynolds et al. 1994). Bosakowski 
and Smith (1992) found that in eastern forests, 
competition between accipiters and buteos is usu- 
ally minimized by a difference in prey selection, 
with buteos typically having a higher proportion of 
mammals in their diet, and accipiters more avian 
prey. 

Red-tailed Hawks frequently exhibit switching 
behavior, which is the capability to utilize which- 
ever species is most abundant at the time (Steen- 
hof and Kochert 1988). Predators that are gener- 
alists often have weak and variable prey 
preferences and will exhibit switching behavior, 
while specialists with strong or consistent prefer- 
ences do not (Murdoch 1969). The wider niche 
breadth of Red-tailed Hawks on the Kaibab Plateau 
indicated weaker preferences compared to the nar- 
rower niche breadth of Northern Goshawks. 

Red-tailed Hawks and Northern Goshawks fre- 
quently occupy similar nesting habitat on the Kai- 
bab Plateau. Both species tend to nest in larger 
trees (x DBH = 68.3-72.5 cm) and mid-slope in 
drainages (x slope position = 0.36-0.37; LaSorte et 
al. 2004) . Because they also nest in close proximity 
(<2000 m apart, USDA Forest Service, North Kai- 
bab Ranger District unpubl. data), we suggest that 
they may also be utilizing the same foraging habi- 
tat. Unfortunately, pellets and prey remains do not 


provide insight into any spatial or temporal parti- 
tioning of prey by these two species. However, our 
research clearly shows Red-tailed Hawks are prey- 
ing upon many of the same species utilized by 
Northern Goshawks. Thus, we believe the potential 
for competition exists. We further hypothesize that 
the more generalist nature of Red-tailed Hawk diet 
and nesting habitat, in combination with the de- 
terioration of late-successional forest habitat and 
concurrent creation of openings on the Kaibab 
Plateau, may tend to exacerbate any potential con- 
flict. We suggest that additional research should be 
implemented to examine this potential competi- 
tion, better quantify numbers of prey in both spe- 
cies’ diets, and determine potential effects on such 
competition on Northern Goshawk management 
alternatives. 

Acknowledgments 

We thank R. Reynolds for first conceiving of this pro- 
ject; he, S. Bayard de Volo, S. Salafsky, and his summer 
crews on the Kaibab Plateau provided field and logistical 
support. A. Hales, R. Lopez, and T. Rohmer assisted with 
data collection. R. Baida and P. Beier offered helpful 
comments on study design, as well as the original man- 
uscript. R. King provided invaluable statistical assistance. 
We also thank D. Andersen and S. Lewis for thorough 
and constructive reviews of the manuscript. This research 
was part of the senior author’s Master of Science project 
at Northern Arizona University, funded by U.S. Depart- 
ment of Agriculture Forest Service, Rocky Mountain Re- 
search Station. 

Literature Cited 

Andersen, D.E., S. DeStefano, M.I. Goldstein, K. Titus, 
C. Crocker-Bedford, JJ- Keane, R.G. Anthony, and 
R.N. Rosenfield. 2003. The status of Northern Gos- 
hawks in the western United States. Wildlife Society 
Technical Review 04-1 . The Wildlife Society, Bethesda, 
MD U.S.A. 

Bai.lam, J.M. 1984. The use of soaring by the Red-tailed 
Hawk. Auk 101:519-524. 

Bloom, P.H., G.R. Stewart, and BJ. Walton. 1986. The 
status of the Northern Goshawk in California, 1981— 


444 


Gatto et al. 


VoL. 39, No. 4 


1983. California Department of Fish and Game, Wild- 
life Management Branch Administration Report 85-1, 

Boal, C.W. 1993. Northern Goshawk diets in ponderosa 
pine forests in northern Arizona. M.S. thesis. Univer- 
sity of Arizona, Tucson, AZ U.S.A. 

AND R.W. Mannan. 1994. Northern Goshawk diets 

in ponderosa pine forests on the Kaibab Plateau. Stud. 
Avian Biol. 16:97-102. 

Bosakowski, T. and D.G. Smith. 1992. Comparative diets 
of sympatric nesting raptors in the eastern deciduous 
forest biome. Can.]. Zool. 70:984—999. 

Carey, A.B. 1984. A critical look at the issue of species- 
habitat dependency. Pages 28-33 in Challenges for 
wildlife and fish-the old growth ecosystem in managed 
forests. Proc. 1983 Technical Session of Wildlife and 
Fish Ecology. Working Group of Society of American 
Forests, Washington, DC U.S.A. 

Cooper, C.F. 1960. Changes in vegetation, structure, and 
growth of southwestern ponderosa pine forests since 
white settlement. Ecol. Monogr. 30:129—164. 

Covington, W.W. and M.M. Moore. 1994. Southwestern 
ponderosa forest structure: changes since Euro-Amer- 
ican settlement. J. For. 92:39-47. 

Crocker-Bedford, C.D. 1990. Goshawk reproduction 
and forest management. Wildl. Soc. Bull. 18:262-269. 

Dyer, D.P. 1978. An analysis of dissimilarity using multi- 
ple environmental variables. Ecology 59:117-125. 

Edington, E.S. 1995. Randomization tests, 3rd Ed. Mar- 
cel Dekker, New York, NY U.S.A. 

Gill, F.B. 1990. Ornithology. W.H. Freeman & Co., New 
York, NYU.S.A. 

Greene, H.W. and F.M. Jaksic. 1983. Food-niche relation- 
ships among sympatric predators: effects of level of 
prey identification. 40:151-154. 

Kaufmann, M.R., R.T. Graham, D.A. Boyce, W.H. Moir, 
L. Perry, R.T. Reynolds, R.L. Bassett, P. Mehlhop, 
C.B. Edminster, W.M. Block, and P.S. Corn. 1994. 
An ecological basis for ecosystem management. USDA 
Forest Service Rocky Mountain Forest 8c Range Ex- 
periment Station, Ft. Collins, CO U.S.A. 

Kennedy, P.L. 1991. Reproductive strategies of Northern 
Goshawks and Cooper’s Hawks in north-central New 
Mexico. Ph.D. dissertation, Utah State University, Lo- 
gan, UT U.S.A. 

Krebs, C.J. 1998. Ecological methodology, 2nd Ed. Ben- 
jamin/Cummings, Menlo Park, CA U.S.A. 

LaSorte, F.A., R. William Mannan, R.T. Reynolds, T.G. 
Grubb. 2004. Habitat association of sympatric Red- 
tailed Hawks and Northern Goshawks on the Kaibab 
Plateau./ Wildl. Manage. 68:307-317. 

Levins, R. 1968. Evolution in changing environments: 
some theoretical explorations. Princeton University 
Press, Princeton, NJ U.S.A. 


Lewis, S.B., M.R. Fuller, and K. Titus. 2004. A compar- 
ison of 3 methods for assessing raptor diet during the 
breeding season. Wildl. Soc. Bull. 32:373-385. 

Moore, T.D., L.E. Spence, C.E. Dugnolle, and W.G. 
Hepworth. 1974. Identification of the dorsal guard 
hairs of some mammals of Wyoming. Wyoming Game 
and Fish Department, Cheyenne, WY U.S.A. 

Murdoch, W.W. 1969. Switching in general predators, 
experiments on predator specificity and stability of 
prey population. Ecol. Monogr. 39:335-354. 

Preston, C.R. and R.D. Beane. 1993. Red-tailed Hawk 
The Birds of North America. The Academy of Natural 
Sciences, Philadelphia, PA U.S.A. and The American 
Ornithologists’ Union, Washington, DC U.S.A. 

Rasmussen, D.I. 1941. Biotic communities of Kaibab Pla- 
teau, Arizona. Ecol. Monogr. 11:231-274. 

Reynolds, R.T., R.T. Graham, M.H. Reiser, R.L. Bassett, 
P.L. Kennedy, D.A. Boyce, G. Goodwin, R. Smith, and 
E.L. Fisher. 1992. Management recommendations for 
the Northern Goshawk in the southwestern United 
States. USDA Forest Service Rocky Mountain Forest 
& Range Experiment Station, Ft. Collins, CO U.S.A. 

, S.M. Joy, and D.G. Leslie. 1994. Nest productiv- 
ity, fidelity, and spacing of Northern Goshawks in Ar- 
izona. Stud. Avian Biol. 16:103-113. 

AND E.C. Meslow. 1984. Partitioning of food and 

niche characteristics of coexisting Accipters during 
breeding. Auk 101:761-779. 

Ricklefs, R.E. AND D. Schluter. 1993. Ecological com- 
munities: historical and geographical perspectives. 
University of Chicago Press, Chicago, IL U.S.A. 

Squires, J.R. and R.T. Reynolds. 1997. Northern Gos- 
hawk. The Birds of North America. The Academy of 
Natural Sciences, Philadelphia, PA and The American 
Ornithologists’ Union, Washington, DC U.S.A. 

Stains, H.J. 1958. Field key to guard hair of middle west- 
ern furbearers. J. Wildl. Manage. 22:95-97. 

Steenhof, K. and M.N. Kochert. 1988. Dietary respons- 
es of three raptor species to changing prey densities 
in a natural environment./. Anim. Ecol. 57:37-48. 

Weaver, H. 1951. Fire as an ecological factor in the 
southwestern ponderosa pine forests./. For. 49:93-98. 

White, M.A. and J.T. Vankat. 1993. Middle and high el- 
evation coniferous forest communities of the North 
Rim region of the Grand Canyon National Park, Ari- 
zona, U.S.A. Vegetatio 109:161-174. 

Williams, C.S. 1938. Aids to the identification of mouse 
and shrew hairs with general comments on hair struc- 
ture and hair determination./. Wildl. Manage. 2:239- 
269. 

Received 15 March 2004; accepted 26 March 2005 

Associate Editor: Clint W. Boal 


J Raptor Res. 39(4) :445-453 
© 2005 The Raptor Research Foundation, Inc. 


BAT PREDATION BY LONG-EARED OWLS IN MEDITERRANEAN 
AND TEMPERATE REGIONS OF SOUTHERN EUROPE 

Ana Maria Garcia^ 

IMEDEA (CSIC-UJB), Miquel Marques 21, E-07190 Esporles, Mallorca, Spain 

Francisco Cervera 

C.R.F. Granja de El Saler, Conselleria de Territori i Habitatge, Avda. dels Pinars 106, E-46012, Valencia, Spain 

Alejandro Rodriguez 

Department of Applied Biology, Estadon Bioldgica de Donana, CSIC, Avda. Maria Luisa s/n, E-41013, Sevilla, Spain 

Abstract. — ^We described spatial and temporal variation in bat consumption by Long-eared Owls {Asio 
otus) at a coastal site of eastern Spain and examined the importance of bats in the diet of this raptor 
in nine temperate and 21 Mediterranean localities of southern Europe. In our study site in Spain, bats 
accounted for 2% of prey items, which is the largest percentage so far reported for the species. The 
vast majority of bats were Pipistrellus spp. Bat predation occurred in all seasons, but was significantly 
higher in spring and summer. The temporal pattern of bat predation was unrelated to temporal variation 
in the consumption of rodents, the dominant prey in the diet. Although a consistent increase in bat 
intake only in years of rodent scarcity predicts an aggregation of occurrences over time, bat occurrence 
during 31 successive seasons was not different from a random sequence. Pellets containing bat remains 
originated mainly from one communal roosting site. Bat remains appeared in pellets from five of 16 
nests, accounting for 17% of prey items on average. In southern Europe, bats occurred in 38% of diets 
in the Mediterranean region, while they were absent in diets from adjacent temperate localities. Our 
results suggest that Long-eared Owls prey on bats rarely and opportunistically in Mediterranean sites, 
but also that bat aggregations could be a locally important food source for some individual owls during 
certain periods. 

Key Words: Long-eared Owl, Asio otus; Chiroptera; diet, Mediterranean basin; trophic plasticity. 


PREDACION de MURCIELAGOS POR el BUHO CHICO EN REGIONES TEMPLADAS YMEDITER- 
rAneas del sur de EUROPA 

Resumen. — Describimos la variacion espacial y temporal en el consumo de murcielagos por parte del 
buho Chico Asio otus en una localidad costera del este de Espana, y examinamos la importancia de los 
quiropteros en la dieta de esta rapaz en 30 local! dades del sur de Europa, 9 de clima templado y 21 de 
clima mediterraneo. En nuestra area de estudio, los quiropteros constituyeron el 2% de las presas 
ingeridas, cifra que representa el mayor consumo conocido para la especie. Casi todos los murcielagos 
consumidos fueron a Pipistrellus spp. Su predacion se produjo en todas las estaciones, pero fue signifi- 
cativamente mas alta en primavera y verano. El patron temporal de predacion de murcielagos no estuvo 
relacionado con la variacion temporal en el consumo de roedores, la presa dominante en la dieta. El 
incremento en el consumo de murcielagos solo en anos en los que los roedores son escasos predice 
una agregacion temporal de las apariciones. Sin embargo, la presencia de murcielagos en la dieta a lo 
largo de 31 estaciones sucesivas no difirio de una secuencia aleatoria. La mayor parte de las egagropilas 
que contuvieron murcielagos procedieron del dormidero comunal. Encontramos restos de quiropteros 
en cinco de los 16 nidos muestreados, donde constituyeron en promedio el 17% de las presas. En 
Europa meridional, los murcielagos aparecieron en el 38% de las dietas de la region mediterranea, 
pero en ninguna de las dietas de la region templada adyacente. Nuestros resultados indican que A. otus 
consume murcielagos con baja frecuenda de forma oportupista en la region mediterranea, pero 
tambien sugieren que las agrtipaciohes de quiropteros pueden ser una fuente de alimento localmente 
importante para algunos individuos durante periodos concretos. 

[Traduccion del autor] 


^ Email address: panamel@ono.com 


445 


446 


Garcia et al. 


VoL. 39, No. 4 


Throughout the boreal and temperate regions 
of Europe, Long-eared Owls {Asio otus) prey almost 
exclusively upon microtine rodents (Herrera and 
Hiraldo 1976, Lundberg 1979, Marks et al. 1999). 
As a result, owl numbers decrease locally with de- 
clining vole (Microtus spp.) populations and, at 
larger spatial scales, owls may use nomadic or ir- 
ruptive movements to track peaks in vole abun- 
dance (Lundberg 1979, Korpimaki 1985, Hanski et 
al. 1991). Such a numerical response is a trait of 
specialist predators that may be explained in part 
by low availability of alternative prey (Weber et al. 
2002) and by high predictability of vole population 
peaks in such ecosystems (Korpimaki 1985). How- 
ever, Long-eared Owls seem to depend less on ro- 
dents at lower latitudes (Bertolino et al. 2001), es- 
pecially in Mediterranean regions (Garcia and 
Cervera 2001) which feature lower environmental 
predictability, resulting in strong seasonal and an- 
nual fluctuations in the abundance of rodents and 
other prey (Blondel and Aronson 1999). Indeed 
diet diversification with decreasing latitude may be 
a general pattern in nocturnal raptors (Herrera 
and Hiraldo 1976, Mikkola 1983, Korpimaki and 
Marti 1995) and other predators (Revilla and Pa- 
lomares 2002, Clavero et al. 2003) , suggesting that 
specialization may not always reflect species-specif- 
ic constraints in physiology or morphology, but be- 
havioral flexibility (Futuyma and Moreno 1988, 
Martin et al. 1995). 

Bats have regularly been reported, albeit in small 
amounts, as prey of a variety of diurnal and noc- 
turnal raptors (e.g., Baker 1962, Ruprecht 1979, 
Barclay et al. 1982). Long-eared Owls are no ex- 
ception. For example, in the British Isles, this spe- 
cies was the second most important bat predator 
among raptors (Speakman 1991). In this paper, we 
describe the pattern of bat consumption by Long- 
eared Owls during an 8-yr period in a Mediterra- 
nean site with good habitat for bats in terms of 
roost (buildings) and food availability (insects in 
rice fields and other flooded areas) . We also review 
European studies of Long-eared Owl diet in the 
Mediterranean basin and the adjacent temperate 
zone to examine the geographical pattern of pre- 
dation on bats. We predicted that occurrence of 
bats in the Mediterranean sites would be higher 
than in temperate sites of similar latitude because 
(1) rodent abundance undergoes pronounced sea- 
sonal and annual fluctuations, and owls must 
search for alternative prey and (2) the season of 
bat activity is longer, and bat average abundance 


higher, in warmer Mediterranean environments 
(Avery 1985, Altringham 1996). 

Methods 

We studied food habits of Long-eared Owls in Devesa 
de I’Albufera, one Mediterranean coastal site near Valen- 
cia city, Spain (39°21'N, 0°19'W). The owl habitat is a 
mosaic of pine forest (Pinus halepensis) with dense un- 
derstory and open areas, mostly dunes and mesic inter- 
dune depressions (Costa et al. 1982). This forested land- 
scape is highly disturbed (many buildings and regular 
recreational activities) and surrounded by a large ex- 
panse of rice fields. From November 1995 to June 2003, 
we collected owl pellets from beneath roost and nest sites 
on a monthly basis. We identified prey remains and, for 
each pellet, determined occurrence and minimum num- 
ber of individuals of prey species. Using these data, we 
analyzed spatio-temporal fluctuations in bat predation. 
For analysis of seasonal variation in bat consumption, sea- 
sons were defined as winter ( January-March) , spring 
(April-June) , summer ( July-September) , and fall (Oc- 
tober-December) . 

We carried out the biogeographic comparison of bat 
predation using data from 30 diet studies from southern 
Europe (Table 1). Each study area was assigned to the 
Mediterranean or the temperate climate region accord- 
ing to Emberger et al. (1963; Fig. 1). We excluded north- 
ern temperate localities to avoid diets almost completely 
dominated by voles. For diets containing bats, we used 
Spearman correlation analysis to test the hypothesis that 
proportion of bats decreased with increasing latitude and 
altitude. 

Results and Discussion 

Bat Consumption in Devesa de I’Albufera, Coast- 
al Spain. We collected 2012 pellets that contained 
6210 prey items. Pellets containing bat remains 
originated mainly from a communal roosting site 
(60%) and also near 16 nest sites, where we re- 
corded successful owl reproduction. Bats account- 
ed for 2% of prey items (Table 1), which is the 
largest percentage thus far reported for Long- 
eared Owls (Marti 1976, Mikkola 1983; Table 1). 
Of 126 bats, 124 were pipistrelle bats {Pipistrellus 
spp; Table 2) . Bat remains occurred in all seasons, 
but predation on bats was significantly higher dur- 
ing the peak of bat activity and abundance in 
spring and summer (G = 47.3, df = 3, P < 0.001; 
Fig. 2) . In our study area, the first flights of young 
pipistrelle bats take place between mid-July and 
mid-August (D. Almenar and M. Monsalve pers. 
comm.), and during the initial 2 wk their flight 
skills are less than those of adults (Blanco 1998). 
Thus, the combination of the annual peak in abun- 
dance associated with the emergence of young bats 
and their relatively higher vulnerability, associated 
with their reduced flight capability, may help to 


December 2005 


Bat Predation by Long-eared Owls 


447 


Table 1. Long-eared Owl diet composition in 30 localities of southern Europe. Each locality is assigned to a climatic 
region either Mediterranean or temperate. The percentage of bats, rodents, and other prey are calculated on the 
total number of prey individuals. Numbers assigned to each study area are the same as in Fig. 1. 


Study 

Country 

Climatic 

Region^* 

No. OF 
Prey 

Percent 

Bats 

Percent 

Rodents 

Percent 

Other 

Prey 

Source 

1 

Spain 

M 

874 

0.00 

86.61 

13.26 

Alegre et al. 1989 

2 

Spain 

M 

232 

0.00 

96.10 

3.90 

Delibes et al. 1984 

3 

Spain 

M 

6929 

0.04 

90.70 

9.20 

Araujo et al. 1974 

4 

Spain 

M 

3726 

0.00 

92.60 

7.30 

San Segundo 1988 

5 

Spain 

M 

3185 

0.03 

78.50 

21.50 

Veiga 1980 

6 

Spain 

M 

255 

0.00 

96.50 

2.80 

Lopez-Gordo et al. 1977 

7 

Spain 

M 

804 

0.00 

72.60 

27.40 

Amat and Soriguer 1981 

8 

Spain 

M 

6210 

2.03 

52.77 

45.20 

This study 

9 

Spain 

M 

6249 

0.08 

89.00 

10.70 

Corral et al. 1979 

10 

France 

M 

368 

0.00 

58.40 

39.13 

Kayser and Sadoul 1996 

11 

Italy 

M 

494 

0.00 

90.08 

9.72 

Gerdol and Perco 1977 

12 

Italy 

M 

121 

0.00 

95.87 

4.13 

Gerdol and Perco 1977 

13 

Italy 

M 

103 

0.00 

54.37 

45.63 

Gerdol et al. 1982 

14 

Italy 

M 

1157 

0.00 

93.52 

6.49 

Casini and Magnani 1988 

15 

Italy 

M 

181 

0.00 

97.24 

2.76 

Capizzi et al. 1998 

16 

Italy 

M 

338 

0.30 

98.20 

1.50 

Plini 1986 

17 

Italy 

M 

1787 

0.00 

81.60 

18.50 

Guidoni et al. 1999 

18 

Italy 

M 

201 

0.00 

95.10 

5.00 

Capizzi and Luiselli 1996 

19 

Italy 

M 

— 

0.11 

70.70 

28.40 

Sublimi and Scalera 1991 

20 

Italy 

M 

234 

<1.40 

93.60 

5.00 

Sara 1990 

21 

Greece 

M 

961 

0.30 

87.90 

11.70 

Alivizatos and Goutner 1999 

22 

Italy 

T 

1787 

0.00 

81.60 

18.40 

Bertolino et al. 2001 

23 

Italy 

T 

1836 

0.00 

85.52 

14.44 

Galeotti and Can ova 1994 

24 

Italy 

T 

519 

0.00 

83.63 

16.37 

Mezzavilla 1993 

25 

Italy 

T 

655 

0.00 

93.44 

6.56 

Malavasi 1995 

26 

Italy 

T 

593 

0.00 

90.58 

9.42 

Aloise and Scaravelli 1995 

27 

Italy 

T 

98 

0.00 

79.59 

20.40 

Riga and Capizzi 1999 

28 

Slovenia 

T 

10991 

0.00 

95.48 

4.52 

Tome 2003a 

29 

Romania 

T 

1268 

0.00 

88.18 

11.82 

Marariu et al. 1991 

30 

Switzerland 

T 

4639 

0.00 

99.23 

0.77 

Roulin 1996 


M = Mediterranean and T = temperate region. 


explain the high occurrence of bats in summer 
samples. The relative importance of bats in the diet 
varied by year (G = 79.3, df = 6, P < 0.001; Fig. 
2) . In our study area and other Mediterranean en- 
vironments, rodent populations typically show an- 
nual minima during summer, especially during 
warm years (Blondel and Aronson 1999) . Although 
owl consumption of rodents roughly follows avail- 
ability (Fig. 3) , bat predation was unrelated to the 
proportion of rodents in the summer diet (r^ = 
0.048, N = 8, P = 0.911; Fig. 3). This suggests that 
bats are not specifically sought as an alternative 
prey. The hypothesis that bats are taken only in 
years of marked rodent scarcity also predicts a 
clumped occurrence of bats in the diet over the 8 


study yr. However, bat occurrence during 31 suc- 
cessive seasons did not differ from a random se- 
quence (runs test, Zar 1984; Z = 0.726, P = 0.233; 
Fig. 3). 

Bat remains occurred in 42 pellets (2.1%). In 
27% of these, bats were the only prey item, with 
2-8 individuals per pellet. In 62% of pellets, bats 
occurred together with other prey, but accounted 
for ^50% of prey items in each sample. The dis- 
tribution of the number of individual bats per pel- 
let was aggregated (did not fit a Poisson distribu- 
tion with X = 0.055, = 42.62, df = 1, P< 0.001; 

only pellets with bats, X = 2.643, = 13.23, df — 

3, P = 0.004). These results suggested that bats 
were clumped when captured. Pipistrelle bats are 


448 


Garcia et al. 


VoL. 39, No. 4 



Figure 1. Localities where Long-eared Owl diet was studied in southern Europe. The broken line separates the 
temperate (squares) and the Mediterranean localities (circles) following Emberger et al. (1963). Black circles indicate 
sites where bat predation has been recorded. Numbers correspond to study numbers in Table 1. 


Table 2. Bat species in the diet of Long-eared Owls in the Mediterranean region of Europe. Seasons when predation 
occurred and the range of body mass (g) are also shown. Body masses are from Palomo and Gisbert (2002). Study 
number and location are as in Table 1 and Fig. 1. 


Species 

Body Mass (g) 

Season 

Source 

Study 

Greater horseshoe bat 

14.6-31.6 

Winter 

Alivizatos and Goutner 1999 

21 

(Rhinolophus ferrumequinum) 


All seasons 

This study 

8 

Greater mouse-eared bat 

21.0-35.0 

Winter-Spring 

Corral et al. 1979 

9 

(Myotis myotis) 


Fall 

Sublimi and Scalera 1991 

19 

Lesser mouse-eared bat 

18.0-29.5 

Spring— Summer 

Veiga 1980 

5 

{Myotis blythii) 

Whiskered bat {Myotis mystacinus) 

4.0-8.0 

All seasons 

Plini 1986 

16 

Myotis spp. 


Winter 

Alivizatos and Goutner 1999 

21 

Pipistrellus spp. 

3.5-10.0 

All seasons 

This study 

8 

Serotine bat {Eptesicus serotinus) 

14.0-33.0 

Winter-Spring 

Corral et al. 1979 

9 

Brown long-eared bat 

6.8-12.0 

Spring-Fall 

Araujo et al. 1974 

3 

{Plecotus auritus) 

Schreibers’ bat {Miniopterus 

10.1-20.8 

All seasons 

This study 

8 

schreibersii) 

European free-tailed bat 

22.0-54.0 

Winter 

Alivizatos and Goutner 1999 

21 

{Tadarida teniotis) 


December 2005 


Bat Predation by Long-eared Owls 


449 


(0 

E 

>. 

E 

OL 

(0 

(0 

(0 

4 -* 

CO 

OQ 


o 

A 

E 



Year 


I 1 winter i— i spring summer fall 

Figure 2 . Temporal variation in the number of individuals taken by Long-eared Owls (bars) and in the percent of 
bats in terms of the total number of prey items (line) in Devesa de I’Albufera between 1996 and 2003. 


very abundant in the study area and roost in col- 
onies, often in buildings. Long-eared Owls could 
capture them at emergence as pipistrelle bats leave 
the roosts in large groups, but return as single in- 


dividuals or in small groups and much more 
spaced over time. Although pipistrelle bats do not 
gather while foraging over large rice fields, they 
can form large aggregations when feeding along 


18 n 



FWSS FWSS FWSS FWSSFWSS FWSS FWSS FWS 
1995 1996 1997 1998 1999 2000 2001 2002 2003 


Figure 3. Seasonal variation in the percent (on the total number of prey items) of bats taken by Long-eared Owls in 
Devesa de I’Albufera between 1995 and 2003. Below the bars and for each year, shading intensity of cells indicates the 
ranked percentage of rodents in the diet, from the highest (dark) to the lowest importance (light) in the seasonal diet. 


450 


Garcia et al. 


VoL. 39, No. 4 


drainage channels or near street lamps (Blake et 
al. 1994), and the owls could hunt them there as 
well. Predation by nocturnal raptors on predictable 
accumulations of bats has been previously docu- 
mented (Barclay et al. 1982, Fenton et al. 1994, 
Hoetker and Gobalet 1999). 

The spatial distribution of bat remains during 
the owl breeding season was not random. Bat re- 
mains appeared in pellets from only five (A^ = 205 
prey items) of 16 nests sampled. The mean per- 
centage of bats per positive nest sample was 17.6 
± 13.0% (SD) of prey items. Only in one owl nest 
did bats account for <11% of prey, and the maxi- 
mum observed was 37%. These figures make less 
plausible the idea of an opportunistic capture of 
bats as a result of accidental encounters (Ruprecht 
1979) . Our results were consistent with the hypoth- 
esis of individual differences in ability to catch bats 
or with individual knowledge of the location of pi- 
pistrelle colonies, which may be more profitable to 
exploit than solitary bat species if the emergence 
of large numbers of bats increases hunting success 
(Fenton et al. 1994). Even small-sized pipistrelle 
bats (body mass = 7.5 g) could be a profitable prey 
for Long-eared Owls if available in large quantities. 
The biomass of 8 pipistrelle bats (60 g; the maxi- 
mum number of bats found per pellet) may satisfy 
two thirds of the daily energy needs (93.3 g for a 
280 g owl; Wijnandts 1984), perhaps with little en- 
ergy expenditure during foraging. 

Biogeographic Pattern. We considered 21 diets 
for the Mediterranean region (34 410 prey items) 
and nine diets in the adjacent temperate region 
(22 386 prey items; Fig. 1). In the Mediterranean 
region, 38% of diets included bats as prey, whereas 
bats did not occur in any diet for the temperate 
region (Table 1). These differences in bat occur- 
rence were significant (G = 6.885, df — 1, P = 
0.009). Even excluding our study in coastal Spain, 
where we found an unusually high quantity of bat 
remains, the mean proportion of bats in the Med- 
iterranean diets was significantly higher than in the 
diets of adjacent temperate sites (Mann-Whitney IT- 
test, Z = 1.98, P — 0.048). In the Mediterranean 
region, the overall importance of bats in the diet 
of Long-eared Owls (0.43% of prey-items) was at 
least twice as high as in other geographical areas. 
But the large number of bats in the diet of owls in 
our study area was very influential in this compar- 
ison. In fact, omitting our results from eastern 
Spain, bats only represent 0.06% of prey items 
found in the combined diets from the Mediterra- 


nean region, which is similar to figures found else- 
where. In the diets from North America (23 888 
prey items) and temperate Europe, plus Iraq, re- 
viewed by Marti (1976), bats did not occur. In later 
reviews, Mikkola (1983) and Speakman (1991) 
found that bats accounted for <0.20% of prey 
items in Europe (67 805 prey items) and 0.05% in 
the British Isles (12 870 prey items). 

In the studies we reviewed, bat predation was re- 
stricted to latitudes 37— 43°N and altitudes 0—1400 
m. Differences between localities in bat occurrence 
in the diet could not be attributed to a decline in 
bat species richness northwards, as species richness 
is almost constant at latitudes 35-50°N in Europe, 
which encompass all localities in Figure 1 (Perez- 
Barberia 1991, Mitchell-Jones et al. 1999). Howev- 
er, bat abundance increases with decreasing lati- 
tude and altitude (Perez-Barberia 1991, Kunz and 
Fenton 2003). Moreover, following the temporal 
pattern of insect availability, bats in Mediterranean 
environments show an extended activity season 
(Avery 1985, Altringham 1996, Blondel and Aron- 
son 1999). Indeed, bat predation occurs in all sea- 
sons (Table 2). If an extended period of activity 
and a higher abundance were indicators of in- 
creased availability of bats for owls, we would ex- 
pect increasing bats in the diet with decreasing lat- 
itude and altitude. We found no such correlations 
(latitude, = —0.05, P= 0.91; altitude, = —0.45, 
P = 0.26), but these analyses were based on small 
sample sizes {N = 8 diets containing bats) . 

Long-eared Owls preyed on nine of 29 bat spe- 
cies present in southern Europe (Mitchell-Jones et 
al. 1999; Table 2). Speakman (1991) suggests that 
large bat species would be more profitable prey 
than small ones, and therefore selected by raptors. 
However, Long-eared Owls consumed a variety of 
bat species, very different in body size, and there 
was no bias toward large species (Table 2) . All the 
bat species that owls consumed, except Tadarida te- 
niotis, forage low in open areas (Altringham 1996), 
just as Long-eared Owls do (Mikkola 1983, Tome 
2003b) while hunting terrestrial prey on the wing 
(Marks et al. 1999). Excluding our results in coastal 
Spain, in Mediterranean environments, mean bat 
intake per diet, standardized as bats per 1000 prey 
items, was 3.4 individuals, suggesting that preda- 
tion is in most cases opportunistic (Ruprecht 
1979). Comparable results have been obtained for 
Barn Owls (Tyto alba; Perez-Barberia 1991), which 
are regarded as opportunistic predators of bats. 

We conclude that bat aggregations could be a 


December 2005 


Bat Predation by Long-eared Owls 


451 


locally important food source for some individual 
owls during certain periods, as exemplified by the 
population of Devesa de TAlbufera. More gener- 
ally, this evidence supports the view that Long- 
eared Owls may show substantial trophic plasticity, 
in contrast to their widespread recognition as a ro- 
dent specialist. In other words, their trophic re- 
sponse may be context-dependent rather than im- 
posed by morphological or behavioral constraints 
that typically affect all populations across the range 
of true specialists. 

At the geographical scale, bat abundance does 
not seem to reflect bat availability for Long-eared 
Owls, maybe because hunting strategies for pre- 
ferred prey such as rodents are not compatible 
with a regular exploitation of flying bats. Accord- 
ingly, bats occur in a number of diets across the 
Mediterranean region, but their contribution re- 
mains largely irrelevant. 

Acknowledgments 

We thank D. Almenar, M.A. Monsalve, and J. Quetglas 
for providing helpful information on bat ecology and R. 
Barclay, A, Gastello, D. Tome, and H. Ulmschneider for 
valuable comments that improved the manuscript. A.M. 
Garcia was supported by a grant from the Conselleria de 
Territori i Habitatge, Generalitat Valenciana. A. Rodri- 
guez benefited from a research contract by the Conseje- 
ria de Educacion, Junta de Andalucia. 

Literature Cited 

Alegre, J., A. Hernandez, and AJ. SAnchez. 1989. Datos 
sobre la dieta invernal del buho chico {Asio otus) en 
la provincia de Leon. Donana Acta Vert. 16:305-309. 
Alivizatos, H. and V. Goutner. 1999. Winter diet of the 
Barn Owl ( Tyto alba) and Long-eared Owl {Asia otus) 
in northeastern Greece: a comparison. J. Raptor Res. 
33:160-163. 

Aloise, G. and D. Scaravelli. 1995. Ecologia alimentare 
del gufo comune, Asio otus, in un roost del basso Man- 
tovano. Avocetta 19:110. 

Altringham, J.D. 1996. Bats. Biology and behavior. Ox- 
ford University Press, Oxford, England. 

Amat, J.A. and R.C. Soriguer. 1981. Analyse comparative 
des regimes alimentaires de I’effraire, Tyto alba, et du 
moyen-duc, Asio otus, dans I’ouest de I’Espagne. Alau- 
49:112-120. 

Araujo, J., J.M. Rey, A. Landin, and A. Moreno. 1974. 
Contribucion al estudio del Buho Chico {Asio otus) en 
Espaha. Ardeola 19:397—428. 

Avery, M.I. 1985. Winter activity of pipistrelle bats. J. 
Anim. Ecol. 54:721-738. 

Baker, J.K. 1962. The manner and efficiency of raptor 
depredation on bats. Condor 64:500-503. 

Barclay, R.M.R., C.E. Thomson, and FJ.S. Phelan. 1982. 
Screech Owl, Otus asio, attempting to capture little 


brown bats, Myotis lucifugus, at a colony. Can. Field-Nat 
96:205-206. 

Bertolino, S., E. Ghiberti, and A. Perrone. 2001. Feed- 
ing ecology of the Long-eared Owl {Asio otus) m 
northern Italy: is it a dietary specialist? Can. J. Zool 
79:2192-2198. 

Blake, D., A.M. Hutson, P.A. Racey, J. Rydell, and JR 
Spearman. 1994. Use of lamplit roads by foraging bats 
in southern England. /. Zool. 234:453-462. 

Blanco, J.C. 1998. Mamfferos de Espana I. Insectivoros, 
Quiropteros, Primates y Carnivoros de la peninsula 
Iberica, Baleares y Canarias. Planeta, Barcelona, 
Spain. 

Blondel, J. and J. Aronson. 1999. Biology and wildlife 
of the Mediterranean region. Oxford University Press, 
Oxford, England. 

Capizzi, D., L. Caroli, and P. Varuzza. 1998. Feeding 
habits of sympatric Long-eared Owl, Asio otus. Tawny 
Owl, Strix aluco, and Barn Owl, Tyto alba, in a medi- 
terranean coastal woodland. Acta Ornithol. 33:85-92 

and L. Luiselli. 1996. Feeding relationships and 

competitive interactions between phylogenetically un- 
related predators (owls and snakes). Acta Oecol. 17: 
265-284. 

Casini, L. and a. Magnani. 1988. Alimentazione inver- 
nale di gufo comune, Asio otus, in un area agricola 
dell’Emilia orientale. Avocetta 12:101—106. 

Clavero, M., j. Prenda, and M. Delibes. 2003. Trophic 
diversity of the otter {Lutra lutra L.) in temperate and 
mediterranean freshwater habitats./. Biogeogr. 30:761- 
769. 

Corral, J.F., J. A. Cortes, and J.M. Gil. 1979. Contribu- 
cion al estudio de la alimentacion de Asio otus en el 
sur de Espana. Donana Acta Vert. 6:179-190. 

Costa, M., J.B. Peris, and R. Figuerola. 1982. La vege- 
tacion de la Devesa de FAlbufera. Monografies 01. 
Delegacio del Medi Ambient i Espais Oberts, Ajunta- 
ment de Valencia, Valencia, Spain. 

Delibes, M., P. Brunet-Lecomte, and M. Manez. 1984. 
Datos sobre la alimentacion de la lechuza comun 
{Tyto alba), el buho chico {Asio otus) y el mochuelo 
{Athene noctua) en una misma localidad de Castilla la 
Vieja. Ardeola 30:57-63. 

Emberger, L., H. Gaussen, M. Kassas, and A. Philippis 
1963. Carte bioclimatique de la zone mediterraneen- 
ne (Etude ecologique de la zone mediterrraneenne). 
UNESCO-FAO, Paris, France. 

Fenton, M.B., I.L. Rautenbach, S.E. Smith, G.M. Swa- 
nepoel, J. Grosell, and j. van Jaarsveld. 1994. Rap- 
tors and bats: threats and opportunities. Anim. Behav. 
48:9-16. 

Futuyma, D.J. AND G. Moreno. 1988. The evolution of 
ecological specialization. Annu. Rev. Ecol. Syst. 19:207- 
233. 

Galeotti, P. AND L. Canova. 1994. Winter diet of Long- 
eared Owl {Asio otus) in the Po plain (northern Italy) . 
J. Raptor Res. 28:265-268. 


452 


Garcia et al. 


VoL. 39, No. 4 


Garcia, A.M. and F. Cervera. 2001. Notas sobre la varia- 
cion estacional y geografica de la dieta del Bubo Chi- 
co Asia otus. Ardeola 48:75-80. 

Gerdol, R., E. Mantovani, and F. Perco. 1982. Prelimi- 
nary comparative study on dietary habits of three Stri- 
giform species in the Carso Triestino. Riv. Ital. Ornitol. 
52:55-60. 

and F. Perco. 1977. Ecological observations on 

the Long-eared Owl Asio otus (L) in north-eastern It- 
aly. Boll. Soc. Adriat. Sci. 41:37-59. 

Guidoni, R., D. Capizzi, L. Caroli, and L. Luiselli. 1999. 
Feeding habits of sympatric owls in an agricultural 
and forested landscape of central Italy. Folia ZooL 48: 
199-202. 

Hanski, L, L. Hansson, and H. Henttonen. 1991. Spe- 
cialist predators, generalist predators, and the micro- 
tine rodent cycle. J. Anim. Ecol. 60:353-367. 

Herrera, C.M. and F. Hiraldo. 1976. Food-niche and 
trophic relationships among European owls. Ornis 
Scand. 7:29-41. 

Hoetker, G.M. and K.W. Gobalet. 1999. Predation on 
Mexican free-tailed bats by Burrowing Owls in Cali- 
fornia. Raptor Res. 33:333—335. 

Kayser, Y. and N. Sadoul. 1996. Cas original de preda- 
tion exercee sur des colonies de larides par le hibou 
moyen-duc {Asio otus) dans les salins d’Aigues-Mortes 
(Camargue, France). Nos 43:485-496. 

Korpimaki, E. 1985. Rapid tracking of micro tine popu- 
lations by their avian predators: possible evidence for 
stabilizing predation. Oikos 45:281-284. 

and C.D. Marti. 1995. Geographical trends in tro- 
phic characteristics of mammal-eating and bird-eating 
raptors in Europe and North America. Auk 112:1004- 
1023. 

Kunz, T.H. and B. Fenton. 2003. Bat ecology. Chicago 
University Press, Chicago, IL U.S.A. 

Lopez Gordo, J.L., E. Lazaro, and A. Fernandez Jorge. 
1977. Comparacion de las dietas de Strix aluco, Asio 
otus y Tyto alba en un mismo biotopo de la provincia 
de Madrid. Ardeola 23:189-221. 

Lundberg, a. 1979. Residency, migration and a compro- 
mise: adaptations to nest site scarcity and food spe- 
cialization in three Fennoscandian owl species. Oeco- 
logia 41:273-281. 

Malavasi, D. 1995. Preliminary data on the diet of the 
Long-eared Owl {Asio otus) in an intensively cultivated 
agricultural land of the Po plain near Mantova. Suppl. 
Ric. Biol. Selvaggina 22:255-256. 

Marariu, D., I. Andreescu, and V. Nesterov. 1991. Les 
composants de la nourriture d’hiver d' Asio otus otus 
(L., 1758) du nord-est de Bucarest (Roumanie). Trav. 
Mus. Hist. Nat. “Gngore Antipa” 51 Alb— 420. 

Marks, J.S., RJ. Cannings, and H. Mikkola. 1999. Family 
Strigidae. Pages 76—151 in], del Hoyo, A. Elliott, and 
J. Sargatal [Eds.] , Handbook of the birds of the world. 
Vol. 5. Lynx Editions, Barcelona, Spain. 


Marti, C.D. 1976. A review of prey selection by the Long- 
eared Owl. Condor 78:331-336. 

Martin, R., A. Rodriguez, and M. Delibes. 1995. Local 
feeding specialization by badgers {Meles meles) in a 
mediterranean environment. Oecologia 101:45-50. 

Mezzavilla, E. 1993. Study on the winter diet of the 
Long-eared Owl, Asio otus, in the Treviso province. 
Lav. Soc. Ven. Sc. Nat. 18:173-182. 

Mikkola, H. 1983. Owls of Europe. T. & A.D. Poyser, Cal- 
ton, U.K. 

Mitchell-Jones, A.J., G. Amori, W. Bogdanowicz, B 
Krystufek, P.J.H. Reijnders, F. Spitzenberger, M. 
Stubbe, J.B.M. Thissen, V. Vohralik, and J. Zima. 
1999. The atlas of European mammals. T. and A.D. 
Poyser, London, England. 

Pai.omo, J.L. AND J. Gisbert. 2002. Atlas de los mamiferos 
terrestres de Espaha. Direccion General de Conser- 
vacion de la Naturaleza-SECEM-SECEMU, Madrid, 
Spain. 

Perez-BarberIa, F.J. 1991. Influencia de la variacion lati- 
tudinal en la contribucion de los murcielagos (Chi- 
roptera) a la dieta de la lechuza comun {Tyto alba). 
Ardeola 38:61—68. 

Plini, P. 1986. Primi dati sull’alimentazione del gufo co- 
mune, Asio otus, nel Lazio. Avocetta 10:41-43. 

Reviiia, E. and E. Palomares. 2002. Does local feeding 
specialization exist in Eurasian badgers? Can. J. Zool. 
80:83-93. 

Riga, F. and D. Capizzi. 1999. Dietary habits of the Long- 
eared Owl, Asio otus, in the Italian peninsula. Acta Or- 
nithol. 34:45-51. 

Roulin, a. 1996. La forme fouisseuse du campagnol te- 
rrestre {Arvicola terrestris) : une proie dominante chez 
le hibou moyen-duc {Asio otus). Nos Oiseaux 43:289- 
294. 

Ruprecht, A.L. 1979. Bats (Chiroptera) as constituents 
of the food of Barn Owls {Tyto alba) in Poland. Ibis 
121:489-494. 

San Segundo, C. 1988. Notas sobre la alimentacion del 
Buho Chico {Asio otus) en Avila. Ardeola 35:150-155. 

SarA, M. 1990. Aspetti della nicchia ecologica degli stri- 
giformi in Sicilia. Naturalista Sicil. 14:109-122. 

Speakman, J.R. 1991. The impact of predation by birds 
on bat populations in the British Isles. Mammal Rev. 
21:123-142. 

SuBLiMi, S. AND L. Scalera. 1991. Dati sulla predazione 
di gufo comune {Asio otus) svernanti in una dolina 
pugliese. Suppl. Ric. Biol. Selvaggina 17:119-122. 

Tome, D. 2003a. Functional response of the Long-eared 
Owl {Asio otus) to changing prey numbers: a 20-year 
study. Ornis Fenn. 80:63—70. 

. 2003b. Nest site selection and predation driven 

despotic distribution of breeding Long-eared Owls 
Asio otus. J. Avian Biol. 34:150-154, 

Veiga, J.P. 1980. Alimentacion y relaciones troficas entre 
la Lechuza Comun {Tyto alba) y el Buho Chico {Asio 


December 2005 


Bat Predation by Long-eared Owls 


453 


otus) en la sierra de Guadarrama (Espana). Ardeola'2,b: 
113-142. 

Weber, J.M., S. Aubry, N. Ferrari, C. Fischer, N. Lachat 
Feller, J.S. Meia, and S. Meyer. 2002. Population 
changes of different predators during a water vole cy- 
cle in a central European mountainous habitat. Eco- 
graphy 25:95-101. 


WijNANDTS, H. 1984. Ecological energetics of the Long- 
eared Owl (Asia otus). Ardea 72:1-92- 
Zar, J.H. 1984. Biostatistical analysis, 2nd Ed. Prentice 
Hall, Englewood Cliffs, NJ U.S.A. 

Received 12 August 2004; accepted 3 July 2005 
Associate Editor: James R. Belthoff 


Short Communications 


J. Raptor Res. 39 (4) :454-457 
© 2005 The Raptor Research Foundation, Inc. 


Differential Effectiveness of Playbacks for Little Owls {Athene noctua) 

Surveys Before and After Sunset 


Joan Navarro^ 

Departament de Biologia Animal (Vertebrats) , Facultat de Biologia, Universitat de Barcelona, 

Avda, Diagonal 645, 08028 Barcelona, Spain 

Eduardo Minguez, David Garcia, Carlos Villacorta, Francisco Botella, 
Jose Antonio Sanchez-Zapata, Martina Carrete, and Andres Gimenez 
Area de Ecologia, Departamento de Biologia Aplicada, Universidad Miguel Hernandez, 
Avda de la Universidad s/n, 03202, Elche, Spain 


Key Words: Little Owl, Athene noctua; survey method', 

population surveys', playback effectiveness. 

Most nocturnal owls respond to broadcast of conspe- 
cific recordings, and this technique may be used to study 
their behavior (Galeotti and Pavan 1993, Galeotti et al. 
1997), to map territories (Finck 1990, Lane et al. 2001), 
to identify individuals (Galeotti and Sacchi 2001, Delport 
et al. 2002) , or to study population trends (Exo and Hen- 
nes 1978, Martinez and Zuberogoitia 2004). Several fac- 
tors affecting the effectiveness of playback techniques or 
spontaneous vocalizations have been identified, includ- 
ing response distance (Proudfoot et al. 2002), season 
(Zuberogoitia and Campos 1998), weather (Lengane and 
Slater 2002) , gender, and social status (Appleby and Red- 
path 1997). However, only a few studies have investigated 
systematically how these factors influence playback meth- 
odology (McGarigal and Fraser 1985, Redpath 1994, Cen- 
tili 2001). When comparing the accuracy of sampling us- 
ing spontaneous owl vocalizations or conspecific 
playbacks, several authors demonstrated that sampling 
error increases when using spontaneous calls (McGarigal 
and Fraser 1985, Haug and Didiuk 1993, Redpath 1994). 
However, for Eurasian Eagle-Owl {Bubo bubo), passive au- 
ditory surveys provide better results than surveys based 
on broadcast calls (Penteriani and Pichera 1991, Marti- 
nez and Zuberogoitia 2002). 

The Little Owl {Athene noctua) is a territorial species 
widely distributed in Palearctic regions. This small raptor 
inhabits a wide variety of semi-open areas, from steppes 
and stony semideserts to farmlands and open woodlands, 
and villages and urban areas (Cramp 1985). Little Owls 
prey on insects, small mammals, and birds, and hunt 
both during diurnal and nocturnal hours (Negro et al. 


^ Email address: joannavarro@ub.edu 


1990) . Researchers have surveyed Little Owls by listening 
to spontaneous vocalizations or by playing typical calls to 
provoke the territorial vocalization after sunset (Finck 
1990, Exo 1992, Mastrorilli 1997, Zuberogoitia and Cam- 
pos 1998, Verwaerde et al. 1999, Pirovano and Galeotti 
1999) or before sunset (Martinez and Zuberogoitia 
2004) . Territorial defense is performed mostly by males, 
which are more vocal than females (Mikkola 1983, Finck 
1990, Zuberogoitia and Campos 1998). 

Here, we examine the effectiveness of the playback 
method to detect Little Owls before or after sunset. Spe- 
cifically, we tested whether broadcast before or after sun- 
set would elicit the greater response frequency and 
whether duration of playback affected the Little Owl re- 
sponse rate. We also offer some suggestions to improve 
the survey methodology. 

Study Area and Methods 

We conducted the study in Clot de Galvany Council 
Natural Park (southeastern Spain, Province of Alicante). 
The study area (ca. 650 ha.) was characterized by a mo- 
saic of shrubs, saline grasslands, and mixed forest, inter- 
spersed with extensive abandoned arboreal cultures such 
as almond and olive trees. The area exhibits a semiarid 
Mediterranean climate (Sancho and Lopez 2002). 

Between 19 April-17 May 2002 (the courtship and ter- 
ritorial defense period; Mikkola 1983, Finck 1990), we 
surveyed twice (before and after sunset) for Little Owls 
at 14 permanent stations during five sessions (10 d). 
Each day, survey stations were sampled with three differ- 
ent “survey” experiments, seven before and seven after 
sunset. The first survey was made 2 hr before sunset 
(1900-2100 H). The second survey was done 2 hr after 
sunset (2130-2330 H). These survey experiments were’ 
(1) Spontaneous Calls: the observer listened for 2 min, 
registering the response rate (number of different Little 
Owls heard); (2) First Playback: after the spontaneous 
calls trial, a territorial intrusion was simulated by broad- 
casting territorial calls of Little Owl for 2 min using a 


454 


December 2005 


Short Communications 


455 


cassette-player. Then, the observer listened for 1 min, re- 
cording the response; (3) Second Playback: after the first 
playback (1 min later), a new territorial intrusion was 
simulated by broadcasting territorial calls for 2 min by 
using a cassette-player. The observer listened for 1 min, 
recording the total response rate (response rate of first 
and second playback) . The sequence of stations surveyed 
was reversed every other visit. To reduce the number of 
Litde Owls that could be potentially counted twice, the 
distances between survey stations were at least 500 m 
(Finck 1990, Exo 1992, Centili 2001). To avoid differenc- 
es associated with possible bias in sound direction, we 
always used the same cassette-player (power: 4 watts, Sony 
WN-FX 195, Barcelona, Spain), placed on the ground 
with the loudspeakers directed upwards. To minimize the 
potential of lower response rates (i.e., owls becoming less 
responsive because they habituated to our broadcasts), 
we used territorial male calls from five different individ- 
ual owls (Roche 1996, SEO 1998, Llimona et al. 2002). 
We did not conduct experiments on windy or rainy days. 

Because data were not normally distributed, we used 
nonparametric tests for statistical analysis (Zar 1996). To 
avoid pseudoreplication, we used the mean of response 
rate for each survey station. Statistical analyses were car- 
ried out using the SPSS statistical package (SPSS for Win- 
dows 1999). Two-tailed P-values were used throughout 
and statistical significance was set at P < 0.05. 

Results 

Before sunset, spontaneous Little Owl calls were heard 
at only two stations (Table 1). In contrast, Little Owls 
were detected after sunset by spontaneous calls at nine 
stations (Table 1). As expected, more Little Owls sang 
spontaneously after sunset than before (Kruskal-Wallis 
test, = 10.39, df = 1, P < 0.001). Numbers of Little 
Owls detected by passive auditory surveys were signifi- 
cantly lower than those detected by playback surveys, 
both before and after sunset (Fig. 1). These differences 
were shown in both the comparison with the first play- 
back (before sunset: = 22.07, df = 1, P< 0.001; after 

sunset: x^ = 19.84, df = 1, P < 0.001), and the second 
playback (before sunset: x^ = 28.17, df = 1, P < 0.001; 
after sunset: x^ = 25.73, df = 1, P < 0.001). Playback 
surveys detected more individuals after sunset than be- 
fore (first playback before vs. after sunset: x^ = 6.83, df 
= 1, P < 0.001; second playback before vs. after sunset: 
X^ = 4.51, df = 1, P = 0.03), but there were no differ- 
ences between the two playback experiments within each 
period (first playback vs. second playback before sunset: 
X^ = 0.55, df = 1, P = 0.46, first playback vs. second 
playback after sunset: x^ = 0.03, df = 1, P = 0.87; Fig. 
1 ). 

Discussion 

Our results strongly suggest that nocturnal broadcast 
surveys were the most effective method for surveying Lit- 
tle Owls, both for detecting presence and counting in- 
dividuals or territories (Zuberogoitia and Campos 1998, 
Verwaerde et al. 1999, Centili 2001, van Nieuwenhuyse et 


2.0 1 



Before Sunset After Sunset 


Figure 1. Call rates (calling means for Little Owls per 
survey station) detected without using playback (sponta- 
neous calls) , using first and second playback both before 
and after sunset. 


al. 2002). These results are in accordance with similar 
studies with other owls: Barred Owl {Strix varia; Mc- 
Garigal and Fraser 1985), Tawny Owl {Strix aluco; Red- 
path 1994), Burrowing Owl {Athene cunicularia; Haug and 
Didiuk 1993), Ferruginous Pygmy-Owl {Glauddium brasi- 
lianum; Proudfoot and Beasom 1996), and Long-eared 
Owl {Asio otus; Martinez et al. 2002). 

Interestingly, our study indicated that Little Owls re- 
sponded at a similar rate in the first and the second play- 
back (before and after sunset). The fact that our first 
playback experiment, broadcasting for 2 min, elicited a 
similar number of owls as when both playbacks were in- 
cluded, suggested that 2 min of continuous playback was 
sufficient for detecting Little Owls, which was a shorter 
period than that used in other studies of this species (5 
min, Zuberogoitia and Campos 1998; 4 min, Verwaerde 
et al. 1999; 3 min, Centili 2001). 

Detection rate of Little Owls after sunset was less vari- 
able and higher than before sunset (Table 1, Fig. 1). 
Thus, before sunset, surveys may underestimate the num- 
ber of breeding Little Owls in an area. However, the use 
of playback before sunset would be useful in finding ter- 
ritories and nests because individuals might be observed 
when calling at perches (Martinez and Zuherogoitia 
2004, pers. obs.). 

Estimating the breeding density of owl species is an 
important part of population studies, and comparisons 
are widely used to assess the abundance of a species 
across years and geographical cireas. Biased estimates of 
breeding pair density are misleading and prevent com- 
parisons between studies. Thus, increased efficiency of 
survey methods and knowledge of potential error is nec- 
essary. Our results suggest broadcasting 2 min of conspe- 



Table 1. Mean call rates (number of different owls calling at each survey station /number of trials ±SD) of Little Owls at Clot de Galvany Park, Alicante Province, 
Spain. Data were obtained during five visits/ survey station; protocol included recording spontaneous calls followed by playback of conspecific calls. 


456 


Short Communications 


VoL. 39, No. 4 




















q 


















00 



03 

m 


in 

m 

o 

o 

i-H 

t-H 


in 

CM 




q 

00 

00 

Til 

00 

in 

m 

q 

q 



q 


CJ5 


j 

Cm 

I-H 


d 

d 

d 

d 

d 

d 

t-H 

I-H 

r-H 

d 

1-H 

d 

d 


Q 

+ 1 

+ 1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+ 1 

+1 

+ 1 

+ 1 

+1 

+1 



o 

o 

o 

o 

o 

o 

o 

o 

O 

o 

o 

o 

o 

o 

esn 


0 

0^1 


cy 

CD 

00 

CM 

q 

CD 

(M 

cy 

q 

q 

q 

q 

q 


u 

w 



c6 

d 

d 

GO 

l-H 

iM 

d 

^H 

f-H 

d 

1-H 

d 

1-H 


</3 
















H 

O 
















c/5 



03 


o 


05 




m 

m 

f— H 


C33 

G75 

Z 



CC 

q 

q 

ty 

q 

00 

00 

1-H 

m 



q 

00 


cAi 

§ 


d 

d 

t-H 

d 

d 

d 

d 

1-H 

d 

d 

d 

hH 

d 

d 


+1 

+1 

+ 1 

+ 1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 


H 

o 

o 

o 

o 

o 

o 

05 

o 

o 

o 

o 

o 

O 

o 



c/7 

g 

CD 

q 

q 

q 

q 

q 

cq 

CM 

q 

q 

00 

q 

q 

q 

q 


t-H 

rH 

d 

rH 


CM 

l-H 

CM 

CM 

1-H 

d 

d 

1-H 

d 

1-H 


c/7 

















J 

ID 

IC3 


in 


O 

m 


05 


m 


03 


(35 


CA) 




TfH 


q 

m 

00 

00 


m 


00 


CM 


p 

d 

d 


d 



d 

d 

d 


d 


d 


d 


0 

+1 

+1 

o 

+1 

o 

+1 

+1 

+ 1 

+1 

o 

+1 

o 

+1 

o 

+1 


5 

o 

o 


o 


o 

o 

o 

o 


o 


o 


C35 


H 


CM 


CM 


00 


00 





CD 


CM 


2; 

d 

d 


d 


d 

d 

d 

d 


d 


d 


d 


0 

















(L 

















C/3 

















q 

















I 


03 

(fO 

m 

o 

00 

i-H 

'tf 

o 

^H 

m 

^H 

o 

CM 




00 

00 

00 


o 

q 


00 

o 

L- 

m 


t-H 

cy 

CD 


CM 

d 

d 

d 

d 

d 


d 

d 

d 

d 

d 

^H 

1-H 

^H 

d 


+ 1 

+1 

+1 

+ 1 

+1 

+ 1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+ 1 


z 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 



0 

00 

to 

oy 

CM 

q 

q 

q 

q 

o 

q 

CD 

o 

GM 

q 

GM 


u 

d 

d 

oi 

d 


CM 

1-H 

1-H 

GM 

fH 

d 

d 

f-H 

^H 

1-H 


W 

































H 

















W 

Zl 

U 

o 

05 


in 

o 


m 


Th 

C05 

03 

C4 

in 

03 

GM 



q 

00 


in 

o 

00 


r*H 

00 

00 

00 

cy 

in 

00 

CD 

P 

C/3 



d 

d 

d 

d 

d 

d 

t-H 

d 

d 

d 

1-H 

d 

d 

d 

J 

CM 

+1 

+ 1 

+ 1 

+1 

+1 

+1 

+1 

+ 1 

+ 1 

+ 1 

+ 1 

+ 1 

+1 

+ 1 

+1 

3 


o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

0 
r-r . 

C/7 

q 

CO 

q 

CO 

q 

cy 

00 

q 

q 

q 


q 

CD 


^H 


Pi 


d 

I-H 

d 



d 

r-H 

fH 

^H 

d 

d 

d 

d 

^H 

PP 

& 

















C/7 

















hJ 







m 

m 







t-H 


(ji 

P 















o 








d 

d 







d 


0 

u 

o 

o 

o 

o 

o 

o 

+ 1 

+ 1 

o 

o 

o 

o 

o 

o 

+ 1 








o 

o 







cn 









(M 

cy 







q 


H 

z 







d 

d 







d 


0 

















Oh 

















C/3 

















z 

















g 















3 


H 


CM 

oo 


in 

CO 


00 

03 

o 

t-H 

CM 

cn 



s 










I — 1 

t-H 

1-H 

1-H 




c/3 

















cific songs just after sunset elicits vocal responses effec- 
tively from resident Little Owls. 

Diferente Effectividad del Playback para Censar Mo- 
CHUELO EuROPEO {ATHENE NOCTUA ) ANTES Y DESPUES DEL 
Anochecer 

Resumen. — Los buhos son dificiles de contar pues son 
poco conspicuos, tienen habitos nocturnos y duante el 
dia permanecen perchados en sitios ocultos. La repro- 
duccion de vocalizaciones previamente grabadas ha sido 
considerada un modo eficiente para determinar la pre- 
sencia de estas sigilosas rapaces. En este estudio exami- 
namos la efectividad y precision de la reproduccion de 
vocalizaciones para detectar individuos de la especie Athe- 
ne noctua. Los censos en los que se reprodujeron llama- 
dos espontaneos fueron menos eficientes que aquellos en 
los que se reprodujeron vocalizaciones coespecificas pre- 
grabadas. La tasa de deteccion luego del atardecer fue 
mayor y menos variable que la tasa previa al atardecer. 
Esto sugiere que los censos nocturnos con vocalizaciones 
podrian ser el metodo mas efectivo para detectar la pre- 
sencia de A. noctua y para contar los individuos y terri- 
torios de esta especie. 

[Traduccion del equipo editorial] 


Acknowledgements 

We thank Glen Proudfoot, Vicenzo Penteriani, Dries 
van Niuwenhuyse, and Dwight Smith for comments on 
the previous manuscript. Field work was carried out un- 
der permission and with collaboration of the Elche Coun- 
cil (Alicante). 

Literature Cited 

Appleby, B.M. and S.M. Redpath. 1997. Indicators of 
male quality in the hoots of Tawny Owls (Strix aluco). 
J. Raptor Res. 31:65-70. 

Centili, D. 2001. Playback and Little Owls {Athene noc- 
tua): preliminary results and considerations. Oriolus 
67:84-88. 

Cramp, S. [Ed.]. 1985. The Birds of the Western Palearc- 
tic. Vol. 4. Oxford University Press. Oxford, England. 
Delport, W., A.C. Kemp, and J.W.H. Ferguson. 2002. Vo- 
cal identification of individual African Wood Owls 
{Strix woodfordii): a technique to monitor long-term 
adult turnover and residency. Ibis 144:30-39. 

Exo, K.M. and R. Hennes. 1978. Empfhlungn zur Met- 
hodik von Siedlungsdichte. Untersuchungen am 
Steinkauz {Athene noctua) . 99:137-141. 

. 1992. Population ecology of Little Owls {Athene 

noctua) in Central Europe: a review. Pages 64-75 in 
C.A. Galbraith, I.R, Taylor, and S. Percival [Eds.], The 
ecology and conservation of European owls. Peter- 
borough, Joint Nature Conservation Committee. (UK 
Nature Conservation, No. 5). 

Finck, P. 1990. Seasonal variation of territory size with 
the Little Owl {Athene noctua). Oecologia 83:68-75. 
Galeotti, P. and G. Pavan. 1993. Differential responses 


December 2005 


Short Communications 


457 


of territorial Tawny Owls (Strix aluco) to the hooting 
of neighbours and strangers. Ibis 135:300-,S04. 

AND R. Sacchi. 2001. Turnover of territorial Scops 

Owls {Otus scops) as estimated by spectrographic anal- 
yses of male hoots. /. Avian Biol. 32:256-262. 

, , and E. Perani. 1997. Cooperative de- 
fense and intrasexual aggression in Scops Owls {Otus 
scops): responses to playback of male and female calls. 
J. Raptor Rss. 31:353—357. 

Haug, E.A. AND A.B. Didiuk. 1993. Use of recorded calls 
to detect Burrowing Owls. J. Field Ornithol. 64:188- 
194. 

Lane, W.H., D.E. Andersen, and T.H. Nicholes. 2001. 
Distribution, abundance and habitat use of singing 
male Boreal Owls in northeast Minnesota. J. Raptor 
Res. 35: 130-140. 

Lengane, T. and P.J.B. Slater. 2002. The effects of rain 
on acoustic communication: Tawny Owls have good 
reason for calling less in wet weather. Proc. R Soc. Lon- 
don. B. 269:2121-2125. 

Llimona, R, E. Mateu, and J.C. Roche. 2002. Gma So- 
nora de las Aves de Espaha. 3 CD. Ed. ALOSA, Soni- 
dos de la Naturaleza. Barcelona, Spain. 

Marion, W.R., T.E. O’Meara, and D.S. Maehr. 1981. Use 
of playback recordings in sampling elusive or secretive 
birds. Stud. Avian Biol. 6:81-85. 

Martinez, J.A. and I. Zuberogoitia. 2002. Factors affect- 
ing the vocal behaviour of Eagle Owls {Bubo bubo): 
effects of sex and territorial status. Ardeola 49:1-10. 

AND . 2004. Effects of habitat loss on per- 
ceived and actual abundance of the Little Owl {Athene 
noctua) in eastern Spain. 51:215-219. 

, , J. Colas, and J. Macia. 2002. Use of re- 
corder calls for detecting Long-eared Owls {Asio otus) . 
Ardeola 49:97-102. 

Mastrorilli, M. 1997. Popolazioni di Civetta {Athene noc- 
tua) e selezioni dell’ habitat in un’area di pianura del- 
la provincia di Bergamo. Riv. Mus. Civ. Sci. Nat. 
“E.Caffi” Bergamo 19:15-19. 

McGarigal, K. and J.D. Fraser. 1985. Barred Owl re- 
sponses to recorded vocalizations. Condor 87:552-553. 

Mikkola, H. 1983. Owls of Europe. T. Be A.D. Poyser, 
Berkhamsted, England. 


Negro, J.J., M.J. DE LA Riva, and F. Hiraldo. 1990. Day- 
time activity of Little Owls {Athene noctua) in south- 
western Spain./. Raptor Res. 24:72-74. 

Penteriani, V. AND F. Pinchera. 1991. Effectiveness of the 
acoustic-lure vs. passive auditory surveys in estimating 
the density and distribution of Eagle Owl {Bubo bubo). 
Suppl. Ric. Biol. Selvaggina 16:385-388. 

PiROVANO, A. AND P. Galeotti. 1999. Territorialismo in- 
tra e interspecifico della Civetta {Athene noctua) in 
Provincia di Pavia. Avocetta 23:139. 

Proudfoot, G.A. and S.L. Beasom. 1996. Responsiveness 
of cactus Ferruginous Pygmy-Owls to broadcasted 
conspecific calls. Wild. Soc. Bull. 24:294—297. 

, , F. Cahvez-Ramirez, and J.L. Mays. 2002. 

Response distance of Ferruginous Pygmy-Owls to 
broadcasted conspecific calls. J. Raptor Res. 36:170- 
175. 

Redpath, S.M. 1994. Censusing Tawny Owls {Strix aluco) 
by the use of imitation calls. Bird Stud. 41:192-198. 

Roche, J.-C. 1996. Tous les Oiseaux d’Europe. 4 CD, 296 
especies. 

Sancho, C. and G. Lopez. 2002. Seguimiento de Aves 
Passeriformes en el Paraje Natural del Clot de Gal- 
vany in: SEO-Alicante. Las Aves en Alicante. Anu. Or- 
nitol. Alicante-2000. Alicante, Spain. 

SEO. 1998. Programa Noctua. Seguimiento de Aves Nocturnas 
en Espaha. SEO/Birdlife. 1 CD. 

SPSS FOR Windows. 1999. SPSS Inc., Chicago, IL U.S.A. 

VAN Nieuwenhuyse, D., W. Belis, and S. Bodson. 2002 
Une methode standardisee pour I’inventaire des Che- 
veches d’Athena {Athene noctua). Ales 39:179-190. 

Verwaerde, J., D. VAN Nieuwenhuyse, F. Nollet, and J 
Bracquene. 1999. Satandaardprotocol voor hetinven- 
tariseren van Steenuilen Athene noctua Scop. (Strigidae) 
in de West.Europese Laagvlakte. Oriolus 65:109-116. 

Zar, J.H. 1996. Biostatistical analysis. Prentice Hall Inter- 
national Editions, London, England. 

Zuberogoitia, I. and L.F. Campos. 1998. Censusing owls 
in large areas: a comparison between methods. Ardeo- 
la 45:47-53. 

Received 22 November 2004; accepted 31 August 2005 

Associate Editor: Cheryl R. Dykstra 


458 


Short Communications 


VoL. 39, No. 4 


J. Raptor Res. 39(4):458— 461 
© 2005 The Raptor Research Foundation, Inc. 


King Vultures {Sarcoramphus papa) Forage in Moriche and Cucurit Palm Stands 

Marsha A. Schlee^ 

Museum National d’Histoire Naturelle, Departement Ecologie et Gestion de la Biodiversite, USM 0305, CP 31 Menagerie, 

57 rue Cuvier, 75231 Paris cedex 05, France 


Key Words: King Vulture, Sarcoramphus papa; Mauritia 

flexuosa; Attalea maripa palms, wedge-capped capuchin mon- 
keys-, Cebus olivaceus; foraging association. 

Feeding on palm fruit, particularly drupes of the Af- 
rican oil palm {Elaeis guineensis), has been documented 
for several Old World species of birds of prey (Thiollay 
1978, Barlow 2004). In the New World, fruits of the im- 
ported African oil palm have been consumed by the Tur- 
key Vulture {Cathartes aura ruficollis; Pinto 1965), Yellow- 
headed Caracara (Milvago chimachima; Haverschmidt 
1962), and Black Vulture {Coragyps atratus; Haverschmidt 
1947, Pinto 1965, Elias and Valencia 1982). Both vulture 
species, as well as the Crested Caracara {Caracara cheri- 
way), consume flesh of coconuts {Cocos nucifera; Hav- 
erschmidt 1947, Crafts 1968), and fruits of palms {Maur- 
itia flexuosa and Desmoncus sp.) have been found in 
stomach contents of the Black Caracara {Daptrius ater, 
Haverschmidt 1962). American Black Vultures also feed 
on sweet potatoes (Mcllhenny 1945) and avocado (Rohl 
1949) when carrion is scarce, and Turkey Vultures ingest 
leaves, seeds, and bark of cottonwood trees {Populus 
spp.), apparently as casting material (Davis 1983). No 
published data were found on ingestion of plant matter 
by King Vultures {Sarcoramphus papa), but residents at 
Hato Las Nieves, Venezuela reported that the species 
consumes fruits of the moriche {Mauritia flexuosa) and 
cucurit {Attalea maripa) palms when carrion is scarce (Y. 
Carbonell and A. Mendoza pers. comm.). The observa- 
tions reported in this paper were gathered in an attempt 
to verify these claims. 

Study Area and Methods 

My observations were part of a long-term study (1994- 
2000) on the abundance, population structure, move- 
ment patterns, and foraging strategies of King Vultures 
m the Serranfa de la Cerbatana, Estado Bolivar, Venezue- 
la The study was conducted in the southeastern part of 
the massif at Hato Las Nieves (Sabana Nueva: 6°34'80"N, 
66°12'17'W), a ranch located ca. 125 km south of Caicara 
del Orinoco. The valley of Las Nieves, ca. 20 km long 
(northwest-southeast) and 9 km wide, is dominated by 
lowland shrub savanna, mainly 220-260 m above sea lev- 
el. The bordering mountains reach elevations of 1600 m 
to the west and 1880 m to the north (Cerro de la Cer- 


^ Email address: schlee@mnhn.fr 


batana). The moriche palms {Mauritia flexuosa) can be 
found scattered in the gallery forests or in stands (mor- 
ichales; see Gonzalez Boscan 1987) in the seasonally in- 
undated areas of the valley. The moriche fruits, 3-7 cm 
long, ovate to globular and having an oily mesocarp 
(Borgtoft Pedersen and Balslev 1990), fall to the ground 
when almost ripe and accumulate in the water among 
fallen fronds and debris. Cucurit palms {Attalea maripa = 
Maximiliana regia) occur as stands within the gallery for- 
ests on dry terrain. The fruits, ovate, 5-7 cm long, are 
also rich in oil (Braun 1997) . The observations reported 
here took place during the rainy season, which lasts April 
through October or November. The study periods (24 
June-24 July 1994, 26 June-14 August 1995, 28 May-14 
July 1996, 30 June-18 August 1997) were set up to co- 
incide with the fruiting season reported locally for both 
palm species. Observations were carried out daily, gen- 
erally from 0615-1900 H. The length of each monitoring 
period (up to 5 hr) depended on weather, logistics, and 
whether or not the King Vultures were present in the 
valley. The number of sample days for each study period 
varied from 29-50 d, with the afternoon of arrival and 
morning of departure being counted if observations were 
carried out (Table 1). I used lOX binoculars and ob- 
served from outcrops and other high points. 

Results 

In all four years, moriche and cucurit fruit production 
took place earlier than expected and little remained by 
mid-July. Both palm species suffered from drought in 
1995 and 1996. A mean of 2.1 ± 1.1 (SD) King Vultures 
(range = 1-4) foraged in four different morichales {N 
= 7 occurrences), and 3.4 ±1.0 (range = 2-4) foraged 
{N = 7 occurrences) in two of the cucurit stands, pri- 
marily during the 1994 and 1995 field seasons (Table 1). 
This activity was most often carried out by two adults to- 
gether {N = 5) or three adults and an immature {N = 
6), presumed to be the local birds. Eleven of the 14 oc- 
currences took place when the vultures had not fed on 
livestock carcasses (natural mortality, including jaguar 
\Panthera onca] predation) or inedible parts of slaugh- 
tered animals for 2—3 wk. Eour bouts in the cucurit stands 
(August 1995) followed presumed feeding on the re- 
mains of jaguar-killed native wildlife. 

On 27 June 1994, at 0947 H, four American Black Vul- 
tures flew from the western morichal. As I approached, 
I sighted an adult King Vulture perched low at the edge 
of the palm stand and another adult on the ground near- 


December 2005 


Short Communications 


459 


Table 1. Mean number of King Vultures per study period (1994-97) seen foraging in moriche and cucurit palm 
stands and number of occurrences at Hato Las Nieves, Venezuela. Number of days the vultures fed on livestock 
carcasses and total days of observation are given. 


Number of 

Study King Vultures 

Pebuod X ± SD (range) 


Number of Days 

Frequency ! 

Livestock 

Moriche Cucurit Carcasses Observation 


1994 

2.2 ± 1.0 (1-4) 

3 

3 

15 

29 

1995 

3.3 ± 1.3 (1-4) 

3 

4 

11 

50 

1996 

3.0 ± 0 (3) 

1 

0 

26 

48 

1997 

0 

0 

0 

15 

49 


by at the base of the only fruiting moriche in the area. 
This bird, with head down, just out of sight, seemed to 
be nibbling on something. When both birds flew to near- 
by palms, I could see that their crops were extended. 
Their flight attracted a Turkey Vulture that walked 
around the area but did not feed. At the site, I found 
one unripe moriche fruit with half the mesocarp freshly 
scraped off longitudinally on one side, the marks clearly 
imprinted on the nut, and a large piece of overripe me- 
socarp from another fruit. No debris, no carrion, and no 
live animals were present. I concluded that the King Vul- 
ture had consumed the missing pulp, but this would not 
explain crop extension. The next day after the rains 
stopped (1620 H), I observed an adult sunning in the 
main cucurit stand. As I penetrated into the gallery for- 
est, I came across another King Vulture rummaging in 
the litter at the base of a cucurit palm. The debris con- 
tained no animal matter, only cucurit fruits, whole or rot- 
ting or partially eaten, pieces of mesocarp, and clean ker- 
nels. Several wedged-capped capuchin monkeys {Cebus 
ohvaceus) were in the stand. Two days later (30 June 
1994), two adults were observed in a morichal at ca. 1100 
H and in a cucurit stand in the afternoon, and the fol- 
lowing day three adults and an immature were in the 
same cucurit stand at mid-day. Capuchin monkeys were 
present both days. The last sighting in 1994 took place 
on 7 July, 2 d after the vultures had consumed a dead 
horse. At 0737 H an adult was again perched low at the 
edge of a morichal. It was raining hard (1000 H) when 
1 found the adult foraging under the only fruiting mor- 
iche in the area. The bird flushed upon seeing me, but 
then returned to the morichal again, weaving in and out 
of the vegetation and foraging on the ground as I tracked 
it for ca. 90 min. Moriche fruits were present in the areas 
where the bird had foraged. I found no carrion or live 
animals. The vulture’s crop was extended, indicating it 
had ingested food while in the morichal. 

On 28 June 1995, at 0722 H, two adult King Vultures 
and an American Black Vulture perched at the edge of 
a morichal. An hour later another adult King Vulture and 
an immature joined them, and one of the first adults flew 
to the ground near a palm with fruit. When I investigat- 


ed, 1 found several fruits with the mesocarp wholly or 
partially scraped off. No carrion was present, but a me- 
dium-sized savanna tortoise ( Geochelone carbonaria) was m 
the area and could have been feeding on the fruits ear- 
lier. Later (1157 H), 1 located the immature bird feeding 
at the base of another fruiting moriche. Again, the me- 
socarp on several fruits had been freshly scraped off. No 
debris or carrion was found, and no live animals were 
nearby. 1 concluded that the King Vulture had eaten the 
pulp. Two days later, at 0740 H, an adult perched in the 
gallery forest north of the central camp, and at 0913 H 
I found it foraging on the ground among the few palms 
that remained of a remnant morichal. No carrion was 
present. On 29 July 1995, 8 d after a carcass had been 
consumed, two adults suddenly flew up from one of the 
gallery-forest floors (1645 H) in an area that had fruiting 
moriche, but 1 was unable to investigate further. Then, 
on 5-6 and 9-10 August 1995, three adults and an im- 
mature, presumed to be the same birds as above, were 
sighted in a cucurit stand. Eight Turkey Vultures were 
also present. The King Vultures perched within the up- 
per strata of the canopy of the broad-leafed trees, occa- 
sionally going to the top of the trees to sun or dry, and 
spent the mornings in the stand. A troop of ca. 30 wedge- 
capped capuchin monkeys, with many females transport- 
ing infants, was foraging there on the first two days. The 
area was strewn with large pieces of bark and fallen 
branches. On the last 2 d, fewer capuchins were present, 
but were seen with a pair of red-howler monkeys {Al- 
ouatta seniculus) . 1 found no remains of carrion. 

On 8 June 1996 at 0913 H, 4 d after the vultures had 
eaten livestock carrion, I came upon two adult King Vul- 
tures and an immature resting on a low shrub at the edge 
of a morichal and next to a fruiting moriche palm. Three 
American Black Vultures were feeding on the ground at 
the base of the palm. The crop of one King Vulture was 
slightly extended. I found no carrion and no live animals, 
only palm fruits lacking part of the mesocarp and show- 
ing signs of having been scraped. I concluded that both 
vulture species had been feeding on the fruits. This was 
the only time the King Vultures were observed to forage 
in a morichal in 1996, but very few moriche palms were 


460 


Short Communications 


VoL. 39, No. 4 


fruiting (<10% in the major stands). No foraging took 
place in the cucurit stands, but only two cucurit palms 
had fruit, although fermenting drupes were found at the 
base of others. Livestock carrion was abundant (Table 1), 
and jaguars were in the area. 

The King Vultures were not seen foraging in the palm 
stands in 1997, even though more moriche and cucurit 
were fruiting than in 1996. A small troop of capuchin 
monkeys was present twice in the study area, and domes- 
tic carrion was not often available (Table 1 ) , as most live- 
stock had been removed from the valley. The vultures 
picked over the debris at former carcass sites on 1 2 d and 
followed a jaguar. This feline was known to have come 
through the valley on four occasions, and the King Vul- 
tures presumably fed on the remains of kills on native 
wildlife at or near Las Nieves on 5 d. 

Discussion 

My observations support the claims that the King Vul- 
tures at Hato Las Nieves eat fruits of the moriche palm, 
particularly when carrion is scarce. Eating moriche fruits 
may partially compensate the lack of carrion, 100 g of 
fresh pulp having 10.5 g of fat and 3.0 g of protein, but 
the fruits could have been ingested for their vitamin-min- 
eral content (see Gonzalez Boscan 1987, Borgtoft Ped- 
ersen and Balslev 1990). From the remains of the fruits 
found at the feeding sites, I conclude that the mesocarp 
was scraped off longitudinally, but judging from the size 
of extended crops of some vultures, some fruits, probably 
the smaller ones, may have been swallowed whole. Both 
feeding techniques are used by the Palm-nut Vulture ( Gy- 
pohierax angolensis) when consuming drupes of the Afri- 
can oil and Raphia palms, the ingested kernels being re- 
gurgitated later (Thiollay 1978). Only once in my 
observations was a potential fallen-fruit consumer pre- 
sent — a tortoise. My observations also lend support to 
claims that American Black Vultures eat Mauritia fruits 
(Y. Carbonell and A. Mendoza pers. comm.). Moriche 
fruit-eating appears to be an activity carried out by vul- 
tures local to the study area; however, because north- 
western Bolivar state is one of the few areas in the Ve- 
nezuelan Guayana that has a high concentration of 
morichales (see Gonzalez Boscan 1987), moriche fruit- 
eating by King Vultures could be more widespread than 
at Hato Las Nieves. 

On the other hand, I was not able to confirm that King 
Vultures eat cucurit fruits, although their consumption is 
plausible considering the oil content of the mesocarp 
(Braun 1997). Whenever the King Vultures foraged in 
the cucurit stands, wedge-capped capuchin monkeys 
were also present, which suggests that a foraging associ- 
ation may exist between the two species. The monkey 
troops were also attended by the Turkey Vultures. Raptor- 
monkey associations usually involve opportunistic feeding 
on flushed invertebrates, particularly insects (e.g., Fon- 
taine 1980), and sometimes on vertebrates displaced by 


the movements of monkey troops (e.g., arboreal snakes: 
Zhang and Wang 2000) . 

How could King Vultures benefit from associating with 
Cebus monkeys? Is it only to profit from occasional pri- 
mate mortality? Although 55% of the wedge-capped ca- 
puchin’s diet consists of plant matter, particularly ripe 
fruits, invertebrates are searched out by peeling off loose 
bark, digging into rotting material and sifting through 
leaf debris on the ground (Robinson 1986). Perhaps the 
King Vultures benefit from larvae that are exposed or fall 
to the ground while the troop forages; vultures are 
known to scoop up maggots from decomposing carcasses 
(Houston 1988). Of greater interest is the occasional 
predatory behavior of the capuchins on vertebrates. For 
example, the viscera of lizards may be eaten but the mus- 
cular part left, and the remains of captured frogs dis- 
carded (Robinson 1986). In Cebus capucinus, after wres- 
tling with an Iguana sp., a monkey managed to break off 
30-40 cm of the iguana’s tail, stripped some meat from 
it, and dropped the rest (Baldwin and Baldwin 1977) . By 
following wedge-capped capuchin monkeys, the vultures 
could glean the remains of discarded vertebrate prey. 
However, eating moriche fruit or monitoring monkeys 
did not seem to be as beneficial as picking over the scat- 
tered remains of former carcasses and following jaguars 
to consume the remains of kills. These observations il- 
lustrate some opportunistic feeding strategies used by 
King Vultures. 

SaRCORAMPHUS PAPA FORRAJEA EN MORICHALES Y EN GRUPOS 
DE Palmas de Cucurto 

Resumen. — Este trabajo se realizo durante las epocas llu- 
viosas desde finales de junio de 1994 hasta mediados de 
agosto de 1997 en la Serrania de la Cerbatana, Hato Las 
Nieves, Estado Bolivar, Venezuela. Los datos recolectados 
sostienen las afirmaciones del personal del hato de que 
Sarcoramphus papas come frutos de la palma moriche 
{Mauritia flexuosa), principalmente cuando escasea la ca- 
rrona. Una media de 2.1 individuos de S. papa (rango = 
1-4) forrajearon en los morichales {N = 7 avistamien- 
tos). No pude confirmar las afirmaciones de que S. papa 
come tambien frutos de la palma de cucurito {Attalea 
maripa). Los frutos de ambas especies de palma contie- 
nen aceite. En mis observaciones, una media de 3.4 m- 
dividuos (rango = 2-4) se encontraron vigilando tropas 
del mono capuchino Cebus olivaceus que habian venido a 
forrajear a los rodales de cucuritos. Sugiero que S. papa 
podria asociarse con los monos para aprovechar los m- 
vertebrados que espantan, como las larvas de insectos, 
los restos de vertebrados que capturan y ocasionalmente 
los cadaveres de monos. 

[Traduccion del autor] 

Acknowledgments 

I would like to thank my Venezuelan counterparts, 
Omar Linares (Universidad Simon Bolivar), the late Jose 
Luis Gomez Carredano, and more recently Guilberto 


December 2005 


Short Communications 


461 


Rios (Universidad Nacional Experimental de los Llanos 
Occidentales, Ezequiel Zamora) for their advice and sup- 
port; Ivan de Angelis and the late Yolanda Carbonell for 
allowing this study at Hato Las Nieves; Y. Carbonell and 
Alberto Mendoza for their valuable observations; 
Susanne Renner (Universitat Ludwig Maximilian, Mu- 
nich) for botanical information; and R. Thorstrom, J. 
Bednarz, and an anonymous referee for their helpful 
comments on the manuscript. 

Literature Cited 

Baldwin, J.D. and J.L Baldwin. 1977. Observations on 
Cebus capucinus in southwestern Panama. Primates 18: 
937-941. 

Barlow, C.R. 2004. The utilization of oil-palm kernel by 
Necrosyrtes monachus in The Gambia. Vulture News 51: 
60-62. 

Borgtoft Pedersen, H. and H. Balslev. 1990. Ecuado- 
rian Palms for Agroforestry. AAU Rep. 23:1-123. 
Braun, A. 1997. La Utilidad de las Palmas en Venezuela. 

Fundacion Thomas Merle, Carupano, Venezuela. 
Crafts, R.C., Jr. 1968. Turkey Vultures found to feed on 
coconut. Wibon Bull. 80:327-328. 

Davis, D. 1983. Maintenance and social behavior of roost- 
ing Turkey Vultures. Pages 322-329 in S.R. Wilbur 
and J.A. Jackson [Eds.] , Vulture biology and manage- 
ment. University of California Press, Berkeley, CA 
U.S.A. 

Elias, D.J. and D. Valencia. 1982. Unusual feeding be- 
havior by a population of Black Vultures. Wibon Bull. 
94:214. 


Fontaine, R. 1980. Observations on the foraging associ- 
ation of Double-toothed Kites and white-faced capu- 
chin monkeys. Auk 97:94-98. 

Gonzalez Boscan, V.C. 1987. Los Morichales de los Lla- 
nos Orientales. Ediciones Corpoven, Caracas, Vene- 
zuela. 

Haverschmidt, F. 1947. The Black Vulture and Caracara 
as vegetarians. Connor 49:210. 

. 1962. Notes on the feeding habits and food of 

some hawks in Surinam. Condor 64:154-158. 

Houston, D.C. 1988. Competition for food between 
Neotropical vultures in forest. Ibis 130:402—417. 

McIlhenny, E.A. 1945. An unusual feeding habit of the 
Black Vulture. Auk 62:136—137. 

Pinto, O. 1965. Dos frutos da palmeira Elaeis guineensts 
na dieta de Cathartes aura ruficollis. Hornero 10:276— 
277. 

Robinson, J.G. 1986. Seasonal variation in use of time 
and space by the wedge-capped capuchin monkey, Ce- 
bus olivaceus: implications for foraging theory. Smith- 
son. Contrib. Zool. 431:1-60. 

Rohl, E. 1949. Fauna descriptiva de Venezuela. Bol. Acad 
Cienc. Fis. Mat. Nat. 12:1-495. 

Thiollay, J.-M. 1978. Les rapaces d’une zone de contact 
savane-foret en Cote d’Ivoire: specialisations alimen- 
taires. Alauda 46:147-170. 

Zhang, S. and L. Wang. 2000. Following of brown ca- 
puchin monkeys by White Hawks in French Guiana. 
Coredor 102:198-201. 

Received 19 July 2004; accepted 18 June 2005 


462 


Short Communications 


VoL. 39, No. 4 


J Raptor Res. 39(4):462— 466 
© 2005 The Raptor Research Foundation, Inc. 


Family Break Up, Departure, and Autumn Migration in Europe of a Family 
OF Greater Spotted Eagles {Aquila cij^nga) as Reported by Satellite Telemetry 

Bernd-U. Meyburg^ 

World Working Group on Birds of Prey, Wangenheimstr. 32, D-14193 Berlin, Germany 

Christiane Meyburg 

World Working Group on Birds of Prey, 31 Avenue du Maine, F-73013 Paris, France 

Tadeusz Mizera and Grzegorz Maciorowski 
Agricultural University, Zoology Department, Wojska Polskiego 71 C, 60-625 Poznan, Poland 

Jan Kowalski 

Gugny 2, 19-104 Trzcianne, Poland 


Key Words: Greater Spotted Eagle; Aquila clanga; depar- 

ture; family break up; migration; satellite telemetry. 

The transition of birds of prey to independence is dif- 
ficult to study (Brown and Amadon 1968), as both old 
and young birds stray ever further from the nest site to- 
ward the end of the post-fledging period. We know of no 
study concerning a raptor species in which the departure 
on migration, the break up of the family, and subsequent 
migration have been investigated by satellite tracking. 
Here, we report on a case concerning Greater Spotted 
Eagles {Aquila clanga) . Available information on this spe- 
cies is limited, but Ivanov et al. (1951) and Dementiev 
and Gladkov (1951) believed that Greater Spotted Eagle 
families departed as a unit on migration. 

In the closely related Lesser Spotted Eagle {Aquila po- 
marina), both adults migrated separately, as determined 
by satellite telemetry (Meyburg et al. 2006). The off- 
spring in this species, which were not tracked with satel- 
lite telemetry, normally leave before the parent birds. 
However, in some cases, the females leave before the 
young (Meyburg et al. 2006). Satellite telemetry has so 
far been used to track one adult Greater Spotted Eagle 
(Meyburg et al. 1995a). 

As part of a long-term research program in northeast- 
ern Poland (Mizera et al. 2001), we are endeavoring to 
raise the level of knowledge, and thereby, the protection 
of the Greater Spotted Eagle by making use of the avail- 
able technology (i.e., satellite telemetry) to investigate 
the species’ migration and wintering habits. 

Methods 

In 1996, an entire family of Greater Spotted Eagles was 
fitted with satellite transmitters (Platform Transmitter 


’ Email address: wwgbp@aol.com 


Terminals, PTTs) in the Biebrza river valley of northeast- 
ern Poland. The eagle nest was located in a National Park 
protecting the largest peatlands in Central Europe, in- 
cluding 15547 ha of forests, 18182 ha of agricultural 
land, and 25 494 ha of wetlands — the Biebrza marshes. 
More than 70 natural and semi-natural plant associations 
have been documented in the Biebrza valley. The most 
dominant forest associations include black alder {Alnus 
glutinosa), swampy birch {Betula pubescens) , and peat co- 
niferous forests {Salici-Betuletum) . Frequent anthropogen- 
ic ecosystems found in the valley are pastures, culdvated 
grounds and urbanized areas. One of the greatest threats 
to the park is human modified drainage patterns, which 
causes the invasion of marshes by shrubs and trees. Active 
conservation measures have been applied to stop further 
succession and maintain more natural intermediate suc- 
cessional stages. A broad public awareness campaign is 
in place to encourage the adoption of organic farming, 
as 45% of the park is privately owned. The eagle nest was 
built in dense humid alder {Alnus glutinosa) and birch 
{Betula spp.) forest. 

We used the dho gaza method (Hamerstrom 1963, 
Clark 1981, Bloom 1987) with a Eurasian Eagle Owl 
{Bubo bubo) to trap the adults. By this method, the eagles 
“attacked” the live eagle owl, tethered to a perch and 
got entangled in the dho gaza net. We used transmitters 
supplied by Microwave Telemetry, Inc. (Columbia, MD 
U.S.A.) with a mass of 60 g. They were fitted as back- 
packs, using Teflon ribbon (Bally Ribbon Mills, Bally, PA 
U.S.A.) to attach them to the bird. The young fledgling 
eagle was equipped with a battery-powered transmitter 
with a mass of 60 g and a temperature sensor. To ensure 
as long a life as possible, this radio was programmed to 
operate only at intervals of 4 d and then for only 10 hr. 
The adult birds were htted with solar-powered PTTs 
These were programmed to be in continuous operation, 
provided the level of light was sufficient to generate pow- 
er for the transmitter. 

All location data were analyzed individually and en- 
tered into databases. We used the computer program 


December 2005 


Short Communications 


463 


Mapit (Allison 1997) to plot locations, which were pro- 
vided by Service Argos, Inc. (Toulouse, France), measure 
distances between locations, and trace the migration 
routes. This program is an integrated global mapping 
and digital display system, which computes the great-cir- 
cle distance between one point and another, while dy- 
namically displaying both great-circle and constant-com- 
pass-bearing (rhumb) lines. Great-circle distances are 
physically the shortest distances on a globe. 

Results 

Three eagles in the family broke up when leaving the 
breeding territory. The female was the first to depart (19 
September 1996), at least 2 or 3 d before her young. The 
male was the last to leave (26 September 1996), 1 wk after 
the female. Whereas the adults headed straight for the 
Bosphorus (Fig. 1), the 1687 km covered by the young 
bird terminated in Albania, where this eagle apparently 
perished at the end of October. The young probably left 
on 21 or 22 September 1996, and the male was present 
on 23 September, but only made minimal migration pro- 
gress by midday on 26 September 1996 (Table 1). Due 
to the programming of its transmitter (operation of 4 d) 
the progress of the young could not be determined as 
accurately as that of the two adults. The date of departure 
of the young eagle was assessed from the first locations 
away from the nest, providing an average estimate of 
speed and distance from the breeding territory during 
the first stages of its migration (Table 1,2). On average, 
this eagle flew only 57 km per day. This young eagle may 
have covered the first 257 km up to the first location away 
from the nest in more than 4 or 5 d, and thus, have left 
the breeding area before 21 or 22 September. 

The young bird set off from the breeding territory in 
a southwesterly direction (Fig. 1 ) . The eagle remained in 
Poland until at least 4 October, while the female and 
male were located in the country for the last time on 19 
and 26 September, respectively. The female reached the 
Bosphorus on 14 October and the male on 22 October. 
The young bird apparently met its death in southern Al- 
bania, ca. 70 km south of Tirana and 13 km north of 
Ballesh. All the data from its transmitter (temperature 
and no change of location) after 26 October indicate 
mortality. The transmitter was transmitting signals until 
13 July 1997. However, it is also possible, but much less 
likely, that the eagle lost or removed the transmitter. 

Discussion 

In birds of prey, the transition to independence is dif- 
ficult to study as toward the end of the post-fledging pe- 
riod, both old and young birds stray ever further from 
the nest site. Direct observation does not account ade- 
quately for local movements of raptors as they begin the 
departure process. Nevertheless, a number of studies 
concerning eagles and other raptors (e.g., Alonso et al. 
1987, Morvan and Dobchies 1990, Bahat 1992, Busta- 
mante 1995, Real et al. 1998, Rafanomezantsoa 2000) 
have been published during recent years. 


b 

u 

a 

rb 


<u 


OJ 

OJ 

IZI 

rd 

S 

Oi 

.a 

V 


c/5 

c: 

OJ 


0 

4-1 


4h 

Tio 

cd 

W 

OJ 

+-> 

o 

Oh 

C/5 

4 

(U 

.kJ 

cd 

OJ 

(5 

OJ 

43 

4h 

o 

a; 

4 

0 


Oh 

(U 

■O 

(U 

43 

H 


3 


0 

H 


4/ 

o 

w 

Q 


z: 

o 

H 


w 

§ H 
H ^ 

Q ^ 




H e 

z O 


w 


0, 

w 

Q 


^ ^ 2 

Lh T 

[jj M L_J 

MH l-H 

Pi 


t4 


n ^ 
2 O 

i ^ 

„ - 2 o 

p 52 



Pi 

u 

S 

D 

P 


0 

4 

o 


O 
c« 


cc 

o 


bD 

o 


cj 

43 


1 

43 

2 

3 

o 

u 

xi 

4 

3 

cn 

43 

H 


I— 1 VJ 

(d (d 

CD > 

cd JP 

2 ^ 

T 3 cd O 

■a ^ p 

kH 

(L) 

cn 
lU 
O 


XS CD 
^ O 


Oh 00 


U 

5:1 


0 

o 

cb 

O 


D 

‘y hh — 

^ 

3 -5 pS 


tn 


O 

• ^ 

ct 

u 

bo 


bo V 


V 

CM 

SO flj 

o 43 

4h 

° 'v B 

“ a o 


O 

a 

o 

43 


cd 
43 

o 

0 

■O 03 


(U Oh 
4S t) 
U CCI 

43 ^ 

. c/5 

t 3 cd 

Th 



CD 


Cd X 

CO g 

00 

C5/ 


CD 

00 

CD 


4J 

3 


V 

O 

4 



V 

h/5 

rb' 

V 

u 

1 'H 

T3 


'bc 


3 

V 

> 

-b' 

u 

2 

-HH 

Oh 

V 

qj 

44 

• ^ 

o 

44 


OJ 

C/J 

u 

3 



43 

C/5 


C7i 

d 

OJ 


H 


H 





4 

o 


CM 

a 

o 




Oh 

■4 

Oh 

05 

c /5 

45 

cd 

43 

05 

C /5 

0 

CM 

05 

4 



Ph 



a 

44 

GO 

05 


a 

44 

J> 

ID 

IM 


I> 

r- 


GO 


TfH 

o 

o 

o 

i-H 

o 

o 


4 

4 

cS 

Cd 

Cd 


4 

4 

Oh 

Oh 

O. 

0/ 

0) 

0/ 

c /5 

c /5 

c /5 

CD 

O 

CD 

CM 

CM 

CM 


^ P P 

01 ^ OJ GO 

C/5 ^ C/5 O 
00 


00 

(M 

CD 

05 


2 

3 

a 

ttJ 

t4 


CD 

CM 

CD 

05 


be ^ 

.a ^ 

4 3 

Oh C 
pCG 4 


464 


Short Communications 


VoL. 39, No. 4 



-52°S 


-48°S 


-44°S 


Figure 1. The autumn migration of the three Greater Spotted Eagles in Europe determined by satellite telemetry 
in 1996; dates of arrival at selected points en route are indicated. 


December 2005 


Short Communications 


465 


Table 2. The outward migration of the juvenile Greater Spotted Eagle (see Fig. 1) 1996 was determined by satellite 
telemetry. 


Beginning and 
End of 
Each Stage"* 

Length of the 
Different Stages 
IN km 

Duration 
OF Each 
Stage (days) 

Mean Length 
OF Daily Flight 
Distances 

Countries 

Transversed 

ca 21/22 Sept 
26 Sept: 0044 H 

257 km 

? 

? 

Poland 

26 Sept: 0044 H 
30 Sept: 1134 H 

Roosting 

4.5 

— 

Poland 

30 Sept: 1134 H 
4 Oct: 1534 H 

242 km 

4.5 

54 km/day 

Poland 

4 Oct: 1720 H 
9 Oct: 0005 H 

76 km 

4 

19 km/day 

Poland and Slovakia 

9 Oct: 0547 H 
13 Oct: 0600 H 

329 km 

4 

82 km/day 

Slovakia and Hungary 

13 Oct: 0600 H 
17 Oct: 1327 H 

202 km 

4.5 

45 km/day 

Hungary and Croatia 

17 Oct: 1738 H 
22 Oct: 0109 H 

361 km 

4 

90 km/day 

Bosnia - Herzegovina 

22 Oct: 0109 H 
26 Oct: 0619 H 

220 km 

4 

55 km/day 

Montenegro and Albania 


Each migration stage was 4 d, based on the duty cycle of the satellite transmitter. Location data provided by Argos Service, Inc 
(Toulouse, France). 


Dementiev and Gladkov (1951) and Ivanov et al. 
(1951) believed that Greater Spotted Eagle families de- 
parted together on migration. We know of no study in 
the literature on this species, or any other raptor, in 
which the dates of departure on migration, the break up 
of the family, and their combined or separate migrations 
have been investigated by satellite telemetry. The family 
of Greater Spotted Eagles studied here clearly broke up 
when leaving the breeding territory. 

We also have studied a pair of the closely-related Lesser 
Spotted Eagle using this method over several years. The 
members of this pair migrated separately in 1997—98 and 
1998-99 and overwintered ca. 1000 km apart both years 
in southern Africa (Meyburg et al. 2006). However, in 
this case, the offspring were not tracked. 

Registro de la Ruptura Familia, Partida y Migracion 
DE Otono de Una Familia de Aquilas Moteadas {Aquila 
clanga) en Europa Usando Telemetria Satelital 

Resumen. — ^Ambos adultos y el polluelo de una famila de 
Aquila clanga fueron estudiados mediante telemetria sa- 
tehtal en el noreste de Polonia para determinar la fecha 
de inicio de su migracion, la disolucion de la familia y 
sus patrones de migracion combinados e independientes. 
La familia se disolvio al abandonar el territorio de cria. 
La hembra fue la primera en partir, el inmaduro lo hizo 
alrededor dos o tres dias mas tarde y el macho partio 
una semana despues de la hembra. Los adultos se diri- 
gieron directo al Bosforo. El inmaduro recorrio 1687 km 


hasta Albania, donde aparentemente murio a fines de 
octubre. 

[Traduccion del autor] 

Acknowledgments 

The authors wish to thank the Deutsche Forschungs- 
gemeinschaft (DFG) in Bonn, Germany, for its unstinting 
financial support of the Satellite Telemetry Greater Spot- 
ted Eagle Project, the Polish Environment Ministry in 
Warsaw, and the administration of the Biebrza National 
Park for kindly allowing us to study and, in particular, to 
trap adult birds. We are also grateful to the Poznan Zoo 
for providing a live eagle-owl to use as a decoy in trapping 
the eagles, to Joachim and Hinrich Matthes, as well as 
Mike McGrady, for help in the field, and to Robin Chan- 
cellor for linguistic help. Prof. Kai Graszynski kindly 
made comments on the first draft of the manuscript, as 
well as three anonymous referees. 

Literature Cited 

Allison, J.B. 1997. Mapit. Version 2.0. Allison Software, 
Apollo, PA U.S.A. 

Alonso, J.C., L.M. Gonzalez, B. Heredia, andJ.L. Gon- 
zalez. 1987. Parental care and the transition to in- 
dependence of Spanish Imperial Eagles {Aquila helia- 
cd) in Donana National Park, southwest Spain. Ibis 
129:212-224. 

Bahat, O. 1992. Post-fledging movements of Golden Ea- 
gles {Aquila chrysaetos homeyeri) in the Negev Desert, 
Israel, as determined by radio-telemetry. Pages 612- 


466 


Short Communications 


VoL. 39, No. 4 


621 in LG. Priede and S.M. Swift [Eds.], Wildlife te- 
lemetry: remote monitoring and tracking of animals. 
Ellis Horwood Ltd., New York, NYU.S.A. 

Bloom, P,H. 1987. Capturing and handling raptors. Pag- 
es 99-123 in B.A.G, Pendleton, B.A. Millsap, K.W. 
Cline, and D.M. Bird [Eds.], Raptor management 
techniques manual. National Wildlife Federation, 
Washington, DC U.S.A. 

Brown, L. and D. Amadon. 1968. Eagles, hawks and Fal- 
cons of the world. Vol. 1. Country Life Books, Felt- 
ham, England. 

Bustamante, J. 1995. The duration of the post-fledging 
dependence period of Ospreys {Pandion haliaetus) at 
Loch Garten, Scotland. Bird Study 42:31-36. 

Clark, W.S. 1981. A modified dho-gaza trap for use at a 
raptor banding station. J. Wildl. Manage. 45:1043- 
1044. 

Dementiev, G.P. and N.A. Gladkov. 1951. Birds of the 
Soviet Union. Moscow, Russia. (In Russian). 

Hamerstrom, F. 1963. The use of Great Horned Owls in 
catching Marsh Hawks. Proc. Int. Ornithol. Congr. 13: 
866-869. 

Ivanov, A.I., E.V. Kozlova E.V., L.A. Portenko, and A.Y. 
Tugarinov. 1951. Birds of Soviet Union. Vol. 1. Iz- 
datelstvo Akademi Nauk SSSR, Moscow, Russia. (In 
Russian). 


Meyburg, B.-U., X. Eichaker, C. Meyburg, and P. Pai- 
LLAT. 1995a. Migrations of an adult Spotted Eagle 
tracked by satellite. Brit. Birds 88:357-361. 

, C. Meyburg, and J. Matthes. 2006. Annual cycle, 

timing and speed of migration of a pair of Lesser 
Spotted Eagles {Aquila pomarina) tracked by satellite. 
J. Ornithol. In press. 

Mizera, T, G. Maciorowski, and B.-U. Meyburg. 2001. 
{Aquila clanga (Pallas, 1811) Greater Spotted Eagle] 
Pages 145-148 in Z. Glowadnski [Ed.], Polish red 
data book on animals. Vertebrates. Panstwowe Wydaw- 
nictwo Rolnicze I Lesne, Warszawa, Poland. (In Polish 
with English summary). 

Morvan, R. and F. Dobchies. 1990. Dependance de jeu- 
nes Aigles de Bonelli {Hieraaetus fasciatus) apres I’en- 
vol: variations individuelles. Alauda 58:150-162. 

Rafanomezantsoa, S.A. 2000. Behavior and range move- 
ments during the post-fledging dependence period of 
the Madagascar Fish-Eagle {Haliaeetus vociferoides) . 
Pages 113-119 inR.D. Chancellor and B.-U. Meyburg 
[Eds.], Raptors at risk. Hancock House and WWGBP, 
Berlin, Germany. 

Real, J., S. Manosa, and J. Godina. 1998. Post-nestling 
dependence period in the Bonelli’s Eagle {Hieraaetus 
fasciatus). Ornis Fennica 75:129-137. 

Received 24 November 2003; accepted 5 September 2005 


J Raptor Res. 39(4):466-47l 
© 2005 The Raptor Research Foundation, Inc. 


Seasonal Patterns of Common Buzzard {Buteo buteo) Reiatwe Abundance and Behavior in 

PoLLiNO National Park, Italy 

Massimo Pandolfi, Alessandro Tanferna, and Giorgia GaibanP 
Istituto di Zoologia, Universitd di Urbino, via Oddi 21, 61029 Urbino, Italy 


Key Words: Common Buzzard', Buteo buteo; relative abun- 

dance] roadside surveys. 

Nest-site selection and habitat use have been described 
m the Common Buzzard {Buteo buteo) by several authors 
(e.g., Penteriani and Faivre 1997, Kruger 2002, Lohmus 
2003, Bustamante and Seoane 2004, Sergio et al. 2005), 
but few studies have documented annual variations in the 
abundance and habitat associations of this species (Meu- 
nier et al. 2000). 

We conducted monthly roadside surveys of Common 
Buzzards in a mountainous area of southern Italy. Al- 


^ Present address and corresponding author: Museo di 
Storia Naturale, Dipartimento di Biologia Evolutiva e 
Funzionale, Universita di Parma, Via Farini 90, 43100 
Parma, Italy; e-mail address: gaibani@biol.unipr.it 


though roadside surveys have well-known limitations 
(e.g., Andersen et al. 1985, Fuller and Mosher 1987, Mill- 
sap and LeFranc 1988, Vinuela 1997), they remain a use- 
ful technique for monitoring local abundance and distri- 
bution of raptors (Fuller and Mosher 1987, Ellis et al. 
1990). Because roadside surveys are easy to conduct, they 
can be carried out at frequent intervals. Here, we present 
results from monthly roadside surveys of Common Buz- 
zards. Using these data, we examine habitat associations, 
describe seasonal patterns of Common Buzzard behavior 
and abundance and, in particular, discuss the effective- 
ness of roadside surveys to monitor changes in abun- 
dance. 

Methods 

The Common Buzzard (hereafter buzzard) surveys were 
conducted from October 2000-September 2001 in Polhno 


December 2005 


Short Communications 


467 


transects 

I — I 


Pollino National Park 







% 

§ 

I 

% 

I 


0,0 4,0 8,0 12,0 16,0 20,0 km 



Figure 1. Locations of seven routes (thick dark lines) used for roadside surveys in Pollino National Park, southern 
Italy, in 2000-01 (thin lines indicate the boundaries). 


National Park (39°58'N, 16°08'E), a 1821 km^ area located 
in the southern Italian Apennines (Fig. 1). The elevation 
ranges from 170-2266 m. The land uses include farmlands 
and oak woods {Quercus ilex, Q. pubescens, Q. cenris) in the 
northern section of the park and grassland and beech 
woods {Fagus sylvatica) in the southern portion. Over the 
study period, the mean monthly temperature was 14.5°C, 
with a mean of 28.8°C during July— September and 9.7°C 
during October-February. The annual rainfall in the 12 
mo of the survey was 841 mm. 

We surveyed buzzards along seven paved roads (Fig. 1) 
selected randomly with restrictions (Caughley and Sin- 
clair 1994). Specifically, we rejected routes adjoining 
those roads previously chosen. Each road was surveyed 
once each mo, during the third or fourth wk of the mo 


and only on calm and clear days. We did not sample on 
days with snow, rain, fog, or strong winds. Each month, 
the routes were surveyed over 4 d by means of two cars, 
each one with a driver and two observers. All surveys 
were conducted in the morning (0900-1200 H), typically 
the best time to count raptors (Robbins 1981). We drove 
at a speed of 40—45 km/hr and stopped the car to con- 
firm each sighting. For each buzzard detected, we re- 
corded if it was flying or perched and if it was alone or 
with other buzzards. Also, we discounted any buzzard that 
may have represented a re-sighted bird. 

We calculated the relative abundance as the number 
of buzzards seen per 100 km sampled. For abundance 
computations we excluded the stretches of roads lined by 
trees within tunnels, forests, or villages. Therefore, al- 


468 


Short Communications 


VoL. 39, No. 4 


Table 1. Number of Common Buzzards recorded along 
seven routes in Pollino National Park (southern Italy, 
2000-01). For computation of relative abundance, we 
used the survey length obtained discounting the stretch- 
es of roads lined by trees or passing through tunnels, 
forests, or villages. 


Routes 

Length 
( km) OF 

Routes 

Survey 
Length 
( km) OF 
Routes 

Monthly Mean 
±SD OF 

Relative Abundance 

1 

57.8 

40.4 

11.2 ± 1.8 

2 

18.3 

18.3 

10.2 ± 2.0 

3 

44.4 

28.9 

6.7 ± 2.0 

4 

47.3 

34.0 

12.7 ± 2.7 

5 

38.5 

38.3 

7.7 ± 1.6 

6 

43.8 

24.2 

20.1 ± 5.7 

7 

65.1 

53.6 

9.3 ± 1.2 


though the total road length was 315.2 km, we consid- 
ered for analysis only 237.6 km (survey km in Table 1). 

For the analysis of buzzard habitat associations, we re- 
ported the sightings as presence/absence in a 1 X 1 km 
UTM grid. We created a 1 km buffer on both sides of 
each route and we considered only buzzards observed 
inside this buffer. We analyzed the habitat associations 
within this buffer using the Corine Land Cover 1:100 000 
digital map (Legend level 3, Ministero dell’Ambiente e 
del Terri to rio — Ente Parco), identifying 10 land cover 
types which we then pooled into four general vegetation 
cover types: (1) arable land (cultivated areas regularly 
plowed and generally under a rotation system), (2) het- 
erogeneous agricultural areas (areas principally occupied 
by agriculture, interspersed with natural areas), (3) for- 
ests, and (4) shrub or herbaceous vegetation (Table 2). 
In each 1X1 km grid cells or portions of a cell included 
inside the buffer, we calculated the surface of each cover 
type by means of a Geographical Information System 
(GIS) analysis (Geomedia Professional 2002). Buzzard 
habitat associations were analyzed in four periods: Feb- 
ruary-April (courtship), May-July (incubation and nest- 
ing), August-September (post-fledging), and October- 
January (winter). Although observers recorded the 
number of individual buzzards sighted, we only consid- 
ered their presence/absence in each grid cell. 

We used the Friedman repeated measures analysis of 
variance (F,.) to detect any difference in the relative abun- 
dance among months and among periods. Using the 
same test, we evaluated if the number of flying or 
perched buzzards varied among months or periods. We 
used the Kruskal-Wallis test to detect any difference in 
the relative abundance among routes and the Mann- 
Whitney U test to ascertain whether flying buzzards were 
observed more frequently than perched buzzards. We 
used the arcsin-transformation to convert the proportion 
of sightings composed of buzzard groups. The Kolmo- 
gorov-Smirnov one-sample test (Z) was used to examine 
if the distribution of relative frequencies was uniform. 
Finally, we used a stepwise logistic regression to deter- 


o 


o 

o 

O, 

3 

iso 

"u 

3 

~a 

u 

0 

3 

V 

u 

Cu 

T3 

Sh 

01 
N 
M 

3 

PQ 

3 

O 

s 

6 

o 

U 

t3 

3 


33 

Ki 

33 

3 

OJ 

o 


V 

.3 

Cu 

• rH 

33 

73 

3 

O 

♦ rH 

'o 

(h 

tJ3 

3 

’u 

O 

"a 

X 

<u 


<L) 

•3! 

O 


W 

H 

X 


S 

Cm 

X 


1— I Q3 

t-h CW CD 

I — ( I — i 


03 o e- 
00 m 

d o o 


o 

X 

3 

Q 

w 

o 

eu 


P.H 

X 

w 


O GM OT) 

GO cd d 


CD (D tT 

OO ■'f oo 
d d d 


O 

X 

H 

C/D 


X 

0 

1 

p 

u 

X 


CL, 

X 


m 03 03 
CM CM ^ 


(M O 

00 00 

odd 


y) 

3 

O 

U 


S 

Oh 

X 

W 


04 

30 

d 


03 

(M 


(D J> CD 
00 CD 

d d d 


3 

o 


73 Q 

^ I 

^ § 

0 ^ 

S 

1 


CM 

(L) 


3 

^ I 


cd 

OJ 

d 

•ui 

3 

u 

bo 

cs 

o 

^ bo 

S ’S 
5 Si 


Shrub or herbaceous vegetation 0.83 1.14 0.53 1.60 0.20 3.79 0.80 0.84 

^ P = statistical significance of the Wald statistic, a chi-square distribution used to ascertain if a variable is a significant predictor of the outcome (presence or absence of buzzards; 
Field 2000) . 

Exp (B) = indicator of the change in odds (probability of an event occurring divided by the probability of that event not occurring) resulting from a unit change in the predictor 
(Field 2000). 


December 2005 


Short Communications 


469 



Figure 2. Monthly mean ±SD of the relative abundance of Common Buzzards recorded along seven routes in 
Pollino National Park, southern Italy (2000-01). The dark columns show the relative abundance calculated consid- 
ering both grouped and single individuals; the white columns are the relative abundance calculated based on single 
individuals only. 


mine whether the probability of detecting buzzards var- 
ied among the four cover types in each of the periods. 
Means are presented ±SD. The nonparametric tests were 
from Siegel and Castellan (1988), and the logistic re- 
gression analysis followed Field (2000). All of the statis- 
tical tests were performed with SPSS 10.0 (SPSS 2000). 

Results 

During the study period, we recorded 328 buzzard 
sightings. The mean relative abundance per roadside sur- 
vey was 11.1 ± 4.4 {N = 84) buzzards/ 100 km (Table 1). 
Buzzard abundance varied among months {F^ — 23.1; df 
= 11; P < 0.05; Fig. 2) and, marginally, among periods 
(Fr = 7.6; df = S, P = 0.054). In particular, a post-hoc 
multiple comparisons test showed that the abundance 
was greater during the courtship period than in the other 
three periods and in the incubation-nesting period than 
in the winter (P < 0.05 for all comparisons). However, 
the relative abundance estimates showed a high variation 
for all periods (17.2 ± 11.9 for courtship, 10.2 ± 5.0 for 
incubation and nesting, 7.2 ± 4.3 for post-fledging, and 
9.0 ± 2.4 for winter). 

We found no difference (x^ = 10.3; df = 6; P = 0.11) 
in abundance among routes, although the survey routes 
crossed different land cover types. We detected more 
buzzards flying (87.2%) than perched (12.8%; U = 1.5; 
N = 24; P < 0.0001). The number of perched buzzards 
did not vary among months (p. = 15.6; df = 11; P > 
0.05), whereas the number of flying buzzards did (P^ = 
21.0; df = 11; P< 0,05) (Fig. 3), In addition, the number 
of flying buzzards varied among periods (p. = 10.8; df = 
3; P< 0.05); during courtship, flying buzzards were more 
numerous than in all other periods (Multiple Compari- 
son test P < 0.05). 


Most buzzards detected (67.4%, N = 221) were alone, 
while 23.2% {N = 76) were paired and 9.4% {N = 31) 
were in larger groups, with a mean group size of 3.9 ± 
0.6 individuals. The proportion of sightings composed of 
a group of buzzards showed a uniform distribution 
throughout the year, ranging from 0.4 in July to 0.0 in 
January (Z = 1.5; N= 12; P> 0.05). 

No land cover variable entered the stepwise logistic re- 
gression discriminating between grid cells with or without 
buzzards, in any of the four periods of the year (Table 
2 ). 

Discussion 

The relative abundance of the study population was 
relatively stable throughout the year, apart from a peak 
during the courtship period (February-April) . A possible 
explanation is that during the courtship period, many 
young buzzards return to their natal area. A high ten- 
dency for philopatric movements among dispersers has 
been recorded both in Common Buzzards (Walls and 
Kenward 1998) and other raptors (e.g., Ferrer 1993, 
Newton et al. 1994, Carter 2001). A second explanation 
for the observed increase in relative abundance may be 
linked more to buzzard behavior and detectability than 
to variations in actual population density. During court- 
ship, buzzards participate in more aerial displays than in 
other periods, and are thus more detectable. This expla- 
nation is supported by the higher proportion of flying 
buzzards recorded during the courtship period, particu- 
larly in March, when all the detected individuals were 
flying (Fig. 3). 

The decrease in relative abundance during the incu- 
bation and nesting periods is probably related to the fact 


470 


Short Communications 


VoL. 39, No. 4 


60 


■o 50 

i_ 

CO 

N 

N 

m 40 

T3 

> 

I 30 

O 

o 20 


CO 

E 

3 


10 


B Flying 
□ Perched 


IxJ 



Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sep 


Figure 3. Total number of flying and perched Common Buzzards observed per month during vehicle surveys con- 
ducted at Pollino National Park, southern Italy in 2000-01. 


that half of the breeding population were tending nests 
or were close to their nests most of the time. However, 
contrary to expectations, during the post-fledging peri- 
od, when flying young join the adult population, the rel- 
ative abundance was not higher than that estimated dur- 
ing the other periods. This could be caused by the 
relatively inconspicuous behavior of fledged buzzards 
that often are perched in the immediate neighborhood 
of the nest for several weeks after fledging (Tubbs 1974, 
Tyack et al. 1998). Moreover, because most natal dispers- 
al usually occurs in the autumn (Picozzi and Weir 1976, 
Walls and Kenward 1998), it is possible that our survey 
methods underestimated fledged young during the post- 
fledging period. 

In our study, buzzards did not show a preference for 
any land cover type during any periods of the year. This 
is supported by lack of variation among the seven road- 
side surveys, although these routes crossed different hab- 
itats. This lack of association with cover type may be due 
to the buzzard’s high plasticity and varied diet (Cramp 
and Simmons 1979, Sergio et al. 2002). However, San- 
chez-Zapata and Calvo (1999), Kruger (2002), and Sergio 
et al. (2005) found that buzzard breeding sites were 
linked to some landscape characteristics, particularly for- 
est cover. There are three possible explanations for this 
discrepancy in results. First, the previous studies exam- 
ined the relationship between habitat characteristics and 
nesting sites, while we examined the association of indi- 
vidual locations with land cover type. Second, we record- 
ed buzzards when in flight or perched, whether they were 
hunting or involved in other activities. Finally, the dis- 


agreement could be related to the coarse scale of our 
landscape analysis. However, Sanchez-Zapata and Calvo 
(1999) found a linkage between buzzard nest sites and 
some landscape characteristics using a coarse-scale (1. 
200 000) land use map. 

In summary, our results showed that the relative abun- 
dance of a raptor population recorded by road counts 
may be sensitive to temporal behavioral changes, which 
ultimately affect detectability, thus biasing a potential as- 
.sessment of seasonal variations in numbers. For this rea- 
son, roadside surveys would be a coarse method to mon- 
itor intra-annual population changes, or to compare the 
relative abundance recorded in different years, unless the 
data are collected in the same time period or season. On 
the other hand, we suggest further investigation into 
whether roadside surveys can be a useful tool for long- 
term monitoring of a population when comparing the 
same months in different years. 

Patrones Estacionaits en ia Abundancia Relativa y el 

COMPARTAMIENTO DE BUIEO BUTEO EN EL PARQUE NACION- 
AL POLl.INO, ItAUA 

Resumen. — Se estudio la abundancia relativa de una po- 
blacion de Buteo buteo a lo largo de 12 meses en el Parque 
Nacional Pollino (sur de Italia) para comparar datos en- 
tre meses y para determinar las asociaciones de habitat 
Realizamos censos mensuales en carreteras a lo largo de 
siete rutas (total = 315 km). La media anual de aves 
detectadas fue de 11.1 ± 4.4 individuos/100 km, aunque 
este valor vario significativamente entre meses. Esta varia- 


December 2005 


Short Communications 


471 


cion probablemente reflejo la actividad de vuelo de B. 
buteo y no las fluctuaciones en el niimero de individuos 
durante el ano, ya que la mayoria de los registros tuvie- 
ron lugar durante el periodo de cortejo. Con base en 
nuestros resultados, sugerimos que los censos realizados 
para esta especie a lo largo de carreteras son mas efec- 
tivos durante el periodo reproductivo que en otras epo- 
cas, cuando los individuos realizan vuelos elevados con 
menor frecuencia. Finalmente, la presencia y la distri- 
bucion de B. buteo dentro del parque no se asociaron con 
el tipo de cobertura del suelo. 

[Traduccion del equipo editorial] 

Acknowledgments 

We are indebted to Pollino National Park, which pro- 
vided support for this project. We are grateful to P. Perna 
for GIS analysis and to A. Aradis, B. Carelli, E. Giardi- 
nazzo, N. Pantone, L. Paternostro, R. Rotondaro, P. Sto- 
rino, and S. Urso for their help in the field data record- 
ing. We thank N.E. Seavy, F. Sergio, and G. Boano for 
their comments on an early draft of the paper. 

Literature Cited 

Andersen, D.E., O.J. Rongstad, and W.R. Mitton. 1985. 
Line transect analysis of raptor abundance along 
roads. Wildl. Soc. Bull. 13:533-539. 

Bustamante, J. and J. Seoane. 2004. Predicting the dis- 
tribution of four species of raptors (Aves: Acdpitridae) 
in southern Spain: statistical models work better than 
existing maps./. Biogeogr. 31:295-306. 

Carter, I. 2001. The Red Kite. Arlequin Press, Chelms- 
ford, Essex, England. 

Caughley, G. and A.R.E. Sinclair. 1994. Wildlife ecology 
and management. Blackwell Science, Cambridge, 
England. 

Cramp, S. and K.E.L. Simmons [Eds.]. 1980. Handbook 
of the birds of Europe, the Middle East, and North 
Africa. The birds of the western palearctic. Vol. II. 
Hawks to Bustards. Oxford University Press, Oxford, 
England. 

Ellis, D.H., R.L. Glinski, and D.G. Smith. 1990. Raptor 
road survey in South America. J. Raptor Res. 24:98- 
106. 

Ferrer, M. 1993. Juvenile dispersal behavior and natal 
philopatry of a long lived raptor, the Spanish Imperial 
Eagle. Ibis 134:128-133. 

Field, A. 2000. Discovering statistics using SPSS for Win- 
dows. SAGE Publications, London, England. 

Fuller, M.R. and J.A. Mosher. 1987. Raptor survey tech- 
niques. Pages 37—65 in B.A. Pendleton, B.A. Millsap, 
K.W. Cline and D.M. Bird [Eds.], Raptor manage- 
ment techniques manual. Nat. Wildl. Fed., Washing- 
ton, DC U.S.A. 

Geomedia Professional. 2002. Geomedia Professional 
5.0 for Windows. Intergraph Corporation, Huntsville, 
AL U.S.A. 

Kruger, O. 2002. Analysis of nest occupancy and nest 
reproduction in two sympatric raptors: Common Buz- 


zard {Buteo buteo) and Goshawk {Accipiter gentilis) . Eco- 

graphy 25:523-532. 

Lohmus, a. 2003. Are certain habitats better every year? 
A review and a case study on birds of prey. Ecography 
26:545-552. 

Meunier, F.D., C. Vereyden, and P. Jouventin. 2000. Use 
of roadside by diurnal raptors in agricultural land- 
scapes. Biol. Conserv. 92:291-298. 

Millsap, B.A. and M.N. LeFranc, Jr. 1988. Road transect 
counts for raptors: how reliable are they? /. Raptor Res. 
22:8-16. 

Newton, L, RE. Davies, and D. Mo.ss. 1994. Philopatry 
and population growth of Red Kites, {Milvus milvus), 
in Wales. Proc. R. Soc. Load. B 257:317-323. 

Penteriani, V. and B. Fatvre. 1997. Breeding density and 
landscape-level habitat selection of Common Buzzard 
{Buteo buteo) in a mountain area (Abruzzo Appenni- 
nes, Italy). / Raptor Res. 31:208-212- 

PiCOZZI, N. AND D.N. Weir. 1976. Dispersal and causes of 
death in buzzards. Br. Birds 69:193-220. 

Robbins, C.S. 1981. Effect of time of day on bird activity. 
Pages 275-286 in C.J. Ralph and J.M. Scott [Eds.], 
Estimating numbers of terrestrial birds. Stud. Avian 
Biol. 6. 

SAnchez-Zapata, J.A. and J.F. Calvo. 1999. Raptor distri- 
bution in relation to landscape composition in semi- 
arid Mediterranean habitats. / Appl. Ecol. 36:254—262 

Sergio, R, A. Boto, C. Scandolara, and G. Bogliani. 
2002. Density, nest-sites, diet, and productivity of 
Common Buzzards {Buteo buteo) in the Italian pre- 
Alps./. Raptor Res. 36: 24-32. 

, C. Scandolara, L. Marchesi, P. Pedrini, and V. 

Penteriani. 2005. Effect of agro-forestry and land- 
scape changes on Common Buzzards {Buteo buteo) m 
the Alps: implications for conservation. Anim. Conserv. 
7:17-25. 

Siegel, S. and N J. Castellan, Jr. 1988. Nonparametric 
Statistics for the Behavioral Sciences. McGraw-Hill, 
Inc., New York, NYU.S.A. (Italian translation.) 

SPSS Inc. 2000. SPSS 10.0 for Windows: base, profession- 
al statistics and advanced statistics. SPSS Inc., Chicago, 
IL U.S.A. 

Tubbs, C.R. 1974. The Buzzard. David 8c Charles, Lon- 
don, England. 

Tyack, A.J., S.S. Walls, and R.E. Kenward. 1998. Behav- 
ior in the post-nestling dependence period of radio- 
tagged Common Buzzards {Buteo buteo). Ibis 140:58— 
63. 

ViNUELA, J. 1997. Road transects as a large-scale method 
for raptors: the case of the Red Kite {Milvus milvus) 
in Spain. Bird Study 44:155—165. 

Walls, S.S. and R.E. Kenward. 1998. Movements of ra- 
dio-tagged buzzards {Buteo buteo) in early life. Ibis 140’ 
561-568. 

Received 7 October 2004; accepted 5 September 2005 

Associate Editor: Fabrizio Sergio 


472 


Short Communications 


VoL. 39, No. 4 


J Raptor Res. 39(4):472-475 
© 2005 The Raptor Research Foundation, Inc. 


New Nesting Record and Observations of Breeding Peregrine Falcons 

IN Baja California Sur, Mexico 

Aradit Castellanos, 1 Cerafina Arguelles, Federico Salinas-Zavala, and Alfredo Ortega-Rubio 
Centro de Investigaciones Biologicas del Noroeste, S.C. Apdo. Postal No. 128 CP 23000, 

La Paz, Baja California Sur, Mexico 


Keywords: Peregrine Falcon-, Falco peregrinus; nesting re- 

cord-, Baja California-, Mexico. 

The B^a California peninsula has been an area where 
resident Peregrine Falcons {Falco peregrinus) were com- 
mon in the past (Bancroft 1927, Banks 1969). Historical 
records of their presence in the region were published 
by Bryant in 1889 (see Grinnell 1928). Detailed accounts 
of nesting territories for the peninsula and Gulf of Cali- 
fornia islands were made by Banks (1969), Anderson 
(1976), and Porter et al. (1988). According to Banks 
(1969), prior to 1967, 54 known peregrine locations ac- 
counted for approximately 66 nest sites in this region. 
Porter et al. (1988) identified 67 eyries in this area in 
1976-84. As in other parts of the world from the late 
1960s to the early 1980s, the peninsular Peregrine Falcon 
population declined, likely due to the impact of organ- 
ochlorine pesticides (Kiff 1988). However, published data 
show that recovery began by the late 1980s (Porter et al. 
1988, Castellanos et al. 1997). 

Historical nesting territories were mainly located on 
sea cliffs of the western side of the state of Baja Califor- 
nia, from Tijuana to Santa Catarina, and on islands along 
both coasts of the peninsula (Banks 1969, Porter et al. 
1988, Castellanos et al. 1997, Ruiz-Campos and Contrer- 
as-Balderas 2000). A small number of inland territories 
were also known (Banks 1969, Castellanos et al. 1997). 
However, no nesting pair has ever been reported on Ba- 
hia Magdalena region, which is located on the west coast 
of the peninsula (Fig. 1). Here, we provide the first re- 
port of Peregrine Falcons nesting in this area. Our find- 
ing extends the breeding range of this species to an area 
of the Baja California peninsula lacking suitable natural 
nesting sites. We also report on the reproductive output 
of this pair. 

Methods 

We observed the nest from vantage points 100-200 m 
from the metal tower, where the peregrine nest was lo- 
cated. We monitored activities around the nesting area, 
and with the help of binoculars and a spotting scope, 
recorded the behavior and attendance of adults at the 
nest. During each observation period, we registered the 


^ Email address; arcas04@cibnor.mx 


contents in the nest, the birds’ activities, time, the num- 
ber of interactions between birds, and disturbance events 
that occurred. 

Results and Discussion 

On 25 January 2004, we found a pair of Peregrine Fal- 
cons at 25°05'2T'N and 112°04'41"W in Santa Elenita, a 
remote, abandoned industrial port 7 km north of Adolfo 
Lopez Mateos in Bahia Magdalena, Baja California Sur, 
Mexico. Both birds displayed activities (mutual roosting, 
cooperative hunting excursions, courtship flights) sug- 
gesting they were a territorial pair. The same day we 
sighted another single adult Peregrine Falcon 7.5 km 
south of Santa Elenita. After our finding, we visited Santa 
Elenita between February to mid-October to monitor the 
pair’s reproductive status (20 hr 25 min of observation 
in 9 d between 1000-1600 H). The port is a concrete 
platform about 30 m wide by 120 m long located in a 
mangrove estuary. Above this platform there are four 
metal towers and a central crane 40 m tall. The crane 
has a metal horizontal arm about 30 m long oriented 
northeast to southwest (Fig. 2). The peregrine pair nest- 
ed on an old Osprey {Pandion haliaetus) nest at the ex- 
treme southwestern end of the arm (Fig. 2). On the op- 
posite side was another Osprey nest. A small fishing camp 
near the platform was operating daily during the study. 

On 28 February, we saw another single bird (male Per- 
egrine Falcon, by relative size) approaching the nest. 
Both males “fought,” and after the interaction and the 
intruder left, the nesting male copulated with the female. 
On 18 March, we found the female incubating eggs. 
From 8-28 May, three nestlings were observed in the 
nest, and they began to fly in late June. On 22 July, we 
observed the parents and one of the fledglings. The 
adults were still on the territory on 1-4 and 11-15 Oc- 
tober. 

The tower was shared with an Osprey pair during the 
entire study period. However, both pairs showed toler- 
ance despite their proximity. Daily and incidental human 
disturbance at the hase of the towers was relatively in- 
tense (22 events or 1.08 events/hr, including small boats, 
people, cars, and motocross traffic); however, the pere- 
grine nest was not deserted and breeding was successful. 
Peregrine response to disturbance occurred more fre- 
quently when other birds approached to within about 20 


December 2005 


Short Communications 


473 



Figure 1. Current and historical distribution of nesting territories of Peregrine Falcon in Baja California Sur, Mexico 


474 


Short Communications 


VoL. 39, No. 4 



Figure 2. View of Santa Elenita mangrove estuary in Bahia Magdalena, B. C. S., Mexico, and Peregrine Falcon (a) 
and Osprey (b) nesting sites. 


m of the nest. We observed 16 interactions (0.79 events/ 
hr), including eight with Magnificent Frigatebirds {Fre- 
gata magnificens ) , four with Ospreys, three with Common 
Ravens {Corvus corax), and one with a gull {Larus sp.). 
Peregrines made territorial attacks or cacking-calls in all 
interactions. 

Breeding dates and the number of nestlings and fledg- 
lings reared by this p£iir were similar to those reported 
for other areas on the western coast of the peninsula 
(Porter et al. 1988, Castellanos et al. 1997). Previous re- 
ports (Castellanos et al. 1997) also show the number of 
nestlings and fledglings produced by pairs on the western 
side of the peninsula are greater than those along the 
Gulf of California. Reasons for this higher productivity 
are unknown. However, it may be an indication of a 
healthier environment; the west coast of the peninsula is 
a relatively low organochlorine pesticide-polluted area in 
North America (total DDT concentration levels on Os- 
prey eggs between 5-311 ppm lipid basis; Spitzer et al. 
1977). 

The southwest coastline of the peninsula is quite dif- 
ferent from the northwestern coast. The terrain is rela- 
tively flat and covered by sparse desert shrubs. The lack 
of suitable natural nesting sites such as sea cliffs, ledges 
on vegetated slopes, and high trees precludes establish- 


ment of breeding pairs in spite of a variety and abun- 
dance of shore birds and waterfowl. The Santa Elenita 
towers provided an opportunity for this nest-site-limited 
species to breed. 

Future conservation of Peregrine Falcon south of the 
U.S.A. should be focused on protection of the natural 
landscape (Temple 1988). This strategy is on course in 
Mexico. In 1972, the islands in the Gulf of California 
were protected as wildlife refuges. In 1988, the entire 
central west coast, including the small islands and Lagun- 
as Ojo de Liebre and San Ignacio, was declared a bio- 
sphere reserve. These refuges preserve prime Peregrine 
Falcon breeding range with low human impact. 

Nuevo Registro de Anidacion y Observaciones de Hal- 
coNES Peregrines Reproductores en Baja Caufornia 
S uR, Mexico 

Resumen. — Encontramos una pareja reproductiva y dos 
adultos no reproductivos de Falco peregrinus en Bahia 
Magdalena, en la costa suroeste de Baja California Sur, 
Mexico. El nido estaba localizado en una torre metalica 
en un puerto abandonado. La anidacion fue exitosa y 
tres volantones abandonaron el nido. Nuestro hallazgo 
amplia el rango de anidacion conocido en la peninsula. 

[Traduccion de los autores] 


December 2005 


Short Communk:ations 


475 


Acknowledgments 

We thank A. Cozar, A. Ortiz, E. Rivera, and P. Tamez 
for field assistance. Financial support was provided by 
Centro de Investigaciones Biologicas del Noroeste, S. C. 
and Secretaria del Medio Ambientes y de los Recursos 
Naturales-Consejo Nacional de Ciencia y Tecnologia proj- 
ect 2002-GO 1-0844. We are grateful to Jim Watson, Tom 
Cade, and one anonymous reviewer for their comments 
on an earlier draft of this paper. Thanks to Dr. E. Glazier 
for editing the English-language text. 

Literature Cited 

Anderson, D.W. 1976. The Gulf of California, Mexico. 
Pages 270-271 in R.W. Fyfe, S.A. Temple, and T.J. 
Cade [Eds.], The 1975 North American Peregrine 
Falcon survey. Can. Field-Nat. 99:228-273. 

Bancroft, G. 1927. Notes on the breeding coastal and 
insular birds of central Lower California. Condor 29: 
228-273. 

Banks, R.C. 1969. The Peregrine Falcon in B^ya Califor- 
nia and the Gulf of California. Pages 81—91 in JJ- 
Hickey [Ed.], Peregrine Falcon Populations. Univer- 
sity Wisconsin Press, Madison, W1 U.S.A. 
Castellanos, A., F. Jaramillo, F. Salinas, A. Ortega-Ru- 
BIO, AND C. Arguelles. 1997. Peregrine Falcon recov- 
ery along the west central coast of the Baja California 
peninsula, Mexico./, Raptor Res. 31:1-6. 

Grinnell, J. 1928. A distributional summation of the or- 
nithology of Lower California. Univ. Calif. Publ. Zool. 
32:1-300. 


Kief, L.F. 1988. Changes in the status of the peregrine in 
North America: an overview. Pages 123-139 in T 
Cade, J. Fnderson, C. Thelander and C. White [Eds.], 
Peregrine Falcon populations: their management and 
recovery. The Peregrine Fund, Inc., Boise, ID U.S.A. 

Porter, R.D., M. A. Jenkins, M.N. Kirven, D.W. Anderson 
AND J.O. Keith. 1988. Status and reproductive perfor- 
mance of marine peregrines in Baja California and 
the Gulf of California, Mexico. Pages 105-114 in T. 
Cade, J. Fnderson, C. Thelander and C. White [Eds.], 
Peregrine Falcon populations: their management and 
recovery. The Peregrine Fund, Inc., Boise, ID U.S.A. 

Ruiz-Campos, G. AND AJ. Contreras-Balderas. 2000. 
New northern nesting record of the Peregrine Falcon 
in Baja California, Mexico./. Raptor Res. 34:151. 

Spitzer, P.R., R.W. Risebrough, J.W. Grier, and C.R. Sin- 
DElAR, Jr. 1977. Eggshell thickness-pollutant relation- 
ships among North American Ospreys. Pages 13—19 
mJ.C. Ogden [Ed.], Trans. North American Osprey 
research conference. U.S. Nat. Park Serv. Proc. Ser. 2, 
Washington, DC U.S.A. 

Temple, S.A. 1988. Future goals and needs for the man- 
agement and conservation of the Peregrine Falcon. 
Pages 843-848 in T. Cade, J. Fnderson, C. Thelander 
and C. White [Eds.], Peregrine Falcon populations: 
their management and recovery. The Peregrine Fund, 
Inc., Boise, ID U.S.A. 

Received 7 December 2004; accepted 5 September 2005 

Associate Editor: James W. Watson 


Letters 


J. Raptor Res. 39(4):476-477 
© 2005 The Raptor Research Foundation, Inc. 


A Previously Undescribed Vocalization of the Northern Pygmy-Owl 


Vocalizations of the Northern Pygmy-Owl {Glaucidium gnoma) were summarized by Holt and Peterson (2000, The 
Birds of North America, No. 494, Philadelphia, PA U.S.A.). The known repertoire of adult vocalizations consists of: 
the “toot song,” which functions as the primary song for both sexes; the “trill call,” which often accompanies the 
“toot song” and for which a function is not yet known; and the “chitter call,” which accompanies prey deliveries 
and certain interactions between breeding adults. Also, a “chatter call” used during copulation was described by 
Righter (1995, Colo. Field Ornithol. J. 29:21—23). Here, I describe a previously undocumented vocalization of the 
Northern Pygmy-Owl. 

As part of an ongoing study of Northern Pygmy-Owls in northern Montana, I have radio-tracked 11 (between one 
and five annually) owls since 2001. Owls were tracked before, during, and after the nesting season. Periods of 
behavioral observation were intermittently conducted in durations ranging from ca. 15 min to 4 hr. On three occa- 
sions, I observed a “weet” vocalization, which somewhat resembled a single note of the “toot song,” but was slightly 
prolonged with a “screeching” quality and a slight upward bend in pitch. In abruptness and duration, this call 
superficially resembled the sudden, high-pitched alarm calls of some Spermophilus ground squirrels, but was deeper 
in pitch and slightly more hollow sounding. The most similar avian call that I am familiar with is the ksew call of the 
Northern Saw-whet Owl {Aegolius acadicus; Cannings 1993, The Birds of North America, No. 42, Washington DC, 
U.S.A.). However, that call is lower in pitch, descending, and usually repeated in a brief series, while the weet call I 
describe here was a single ascending note. Proudfoot and Johnson (2000:7, The Birds of North America, No. 498, 
Philadelphia, PA U.S.A.) describe an alarm call of the congeneric Ferruginous Pygmy-Owl {Glaucidium brasilianum) 
as “short and sharp with upward inflection, pee weet, repeated at irregular intervals.” Although I am unfamiliar with 
the Ferruginous Pygmy-Owl alarm call, this describes accurately the Northern Pygmy-Owl vocalization I observed, 
except that the former consists of two syllables and the latter only one. 

On 21 July 2003, while observing a Northern Pygmy-Owl family group, I first observed an adult male give the 
''iveet" call. A few seconds later an adult Northern Goshawk {Accipiter gentilis) flew within 50 m of the family group 
and disappeared. On 1 July 2004 at 0936 H, while observing a family group with recently fledged young, I observed 
both the adult male and adult female give the weet call. A Northern Goshawk appeared and perched briefly in the 
same stand as the family group a few seconds after the calls. In that instance, the male gave the call first and was 
followed by the female. The female then repeated the call one more time just before the goshawk entered the stand 
The female’s call was slightly higher pitched to a degree approximately equivalent to the difference in pitch of the 
“toot song” between the .sexes (Holt and Peterson 2000, pers. obs.). At 1003 H, during the same observation period 
on 1 July 2004, a Northern Goshawk flew through an opening adjacent to the stand in which the family group (same 
stand and same family group as the previous observation) was located. Both adults gave weet calls, and both repeated 
the calls after 5—10 sec. In that instance, several of the seven fledglings had been calling actively before the adults 
gave the weet calls. The fledglings’ calls immediately ceased after the weet calls were given, but resumed less than 30 
sec later, at which time the goshawk was no longer visible to me. 1 am not certain whether any of the Northern 
Goshawks in these instances were aware of the owls. Twice the goshawks simply flew past, and the one time a goshawk 
perched nearby, it quickly left the perch when it apparently became aware of my presence. The contexts of these 
observations suggest that the function of the weet call is an alarm call. During a Northern Pygmy-Owl study in 
Washington, A. Giese (unpubl. data) observed what was likely the call described here, and likewise suspected its 
function as an alarm call. The call was given several times during a 1.5 hr period by an adult female in the presence 
of fledged young. The young generally ceased vocalizing after the call was given. Unlike the calls I observed, however, 
that female vocalized in bouts of 3-4 calls at a time. 

Interestingly, I have observed instances in which Northern Pygmy-Owls might have been expected to give alarm 
calls, but did not give the weet call. I observed a domestic dog in close proximity (<5 m) to a group of fledgling and 
adult owls perched low on branches on two occasions in 2004 and heard no calls. When banding young owls or 
climbing nest trees to check nests, I have observed adults give short versions of the toot song, trill calls, and display 
agitated behavior (e.g., rapid tail twitching, perching close to and staring at the human intruder). However, I have 
not heard the weet call I describe here in those situations. On 7 May 2001, I observed a Northern Goshawk pass 


476 


December 2005 


Letters 


477 


within 5 m of a solitary nonbreeding Northern Pygmy-Owl perched with prey. The owl watched the goshawk intently, 
but gave no call. 

If the weet call was indeed an alarm call, these observations suggested that it may be associated specifically with 
avian predators, as the mammalian (dog and human) observations I described did not elicit the call. Additionally, 
the observation of the solitary owl failing to give a weet call suggests that it may be used only when fledglings or 
mated owls are present. Alarm calls are well documented for many strigids (e.g.. Ferruginous Pygmy-Owl, Proudfoot 
and Johnson 2000; Great Gray Owl, Strix nebulosa, Bull and Duncan 1993, The Birds of North America, No. 41, 
Philadelphia, PA U.S.A.; Long-eared Owl Asio otus, Marks et al. 1994, The Birds of North America, No. 133, Phila- 
delphia, PA U.S.A.) and are often given in response to threats to nests and fledged young. However, further study is 
needed to better understand the causes and contexts of this and other Northern Pygmy-Owl vocalizations. 

I am grateful to the many individuals who have volunteered their time to conduct fieldwork during this study. 
Private contributions have helped make this ongoing study possible. The Conservation Research Foundation, Mar- 
mot’s Edge Conservation, the Owl Research Institute, and the Rocky Mountain Ranger District of the Lewis and 
Clark National Forest have generously provided assistance with various parts of this project as well. A. Giese and G 
Proudfoot provided criticism and field observations that improved the manuscript. — Graham G. Frye (e-mail address: 
ggfrye@rmf-inh,org) Rocky Mountain Front Institute of Natural History, P.O. Box 186, Ghoteau, MT 59422 U.SA. 

Received 12 August 2004; accepted 18 June 2005 
Associate Editor: Ian G. Warkentin 


BOOK REVIEW 


J. Raptor Res. 39(4) :478-479 
© 2005 The Raptor Research Foundation, Inc. 


Hawks and Owls of Eastern North America. By 

Donald S. Heintzelman. 2004. Rutgers University 
Press, Piscataway, NJ U.S.A. viii + 203 pp., 5 color 
plates, numerous black and white photos. ISBN 
0-8135-3350-3. Cloth, $29.95.— Donald Heintzel- 
man ’s original purpose with this volume was to pre- 
pare a second edition of his earlier work, Hawks 
and Owls of North America (1979, Universe Books, 
New York, NYU.S.A.). However, this book restricts 
coverage to the raptors of eastern North America, 
which is loosely defined as birds east of the Missis- 
sippi River, except for parts of Minnesota and On- 
tario. The volume begins with six chapters dealing 
with general topics: Introduction to Raptor Ecolo- 
gy, Hawk Migrations, Owl Migrations and Inva- 
sions, Raptor Conservation, Citizen Scientists, and 
Recreational Raptor Watching; and follows with 
eight chapters of raptor species accounts mostly in 
taxonomic order, beginning with Ospreys {Pandion 
halieatus) and ending with Northern Saw-whet Owls 
{Aegolius acadicus). For the most part, the book is 
nicely illustrated with a selection of outstanding 
black and white photos. 

I found the introductory chapter, “An Introduc- 
tion to Raptor Ecology,” to be well out of date. In 
the preface, Heintzelman acknowledges that this 
chapter includes substantial portions of the text 
from his 1979 contribution. Upon reading this 
chapter, it is clear that many of the ideas expressed 
reflect antiquated ecological opinions and specu- 
lations of the 1960s and early 1970s. Statements 
such as. Northern Goshawks {Accipiter gentilis) are 
beneficial to Ruffed Grouse (Bonasa umbellus) pop- 
ulations, that raptors control numbers of prolific 
rodents, and raptors maintain a delicate and effec- 
tive ecological balance between predator and prey, 
without presentation of supporting data or sources 
are made freely throughout this short chapter. 

The subsequent introductory chapters, although 
brief, provided easier reading and were based on 
more-current information. The chapter on Raptor 
Conservation reported a series of interesting and 
relatively-recent anecdotes. However, when discuss- 


ing habitat degradation and loss, Heintzelman em- 
phasizes the “ambitious, long-term restoration” ef- 
fort of the Lehigh Gap Restoration Project, which 
has a goal of restoring 750 acres of woodland on 
the Kittatinny Ridge in Pennsylvania. From a rap- 
tor perspective, restoring ca. 3 km^ of deciduous 
woodland habitat is extremely trivial and will likely 
have little impact on the population of any raptor. 
I am familiar with several other governmental and 
private (e.g., the Nature Conservancy) land acqui- 
sition/habitat restoration programs that involve 
many thousands of hectares that are much more 
likely to have substantial population impacts on 
several species of raptors. I would have liked to 
have seen Heintzelman discuss some of these ma- 
jor habitat restoration efforts, perhaps in addition 
to smaller isolated projects that he has personally 
spear-headed. 

Some minor distractions for me were the provi- 
sion of selective contact information. Although 
most major raptor conservation organizations were 
mentioned in the volume, contact information was 
only provided for a select few. Perhaps, in this day 
and age in which an interested person can quickly 
obtain contact information by googling the name 
of an organization, this is not necessary. But, why 
provide detailed contact information (postal ad- 
dress, e-mail address, phone numbers) for a few 
selected organizations, and no information for oth- 
ers? 

Probably more bothersome for a professional or- 
nithologist using this volume is the style of not cit- 
ing references in the text. Although most chapters 
are reasonably-well researched and supported with 
references, albeit selectively, all references are in- 
cluded in the back of the volume listed alphabeti- 
cally by chapter. Thus, when you encounter a state- 
ment in a given chapter, identifying the 
responsible source is exceedingly difficult. Also, 
the supporting sources represent a very mixed bag 
in which some very current and important studies 
are mentioned and cited, while most of the refer- 
ences are from state journals and generally repre- 
sent novel anecdotes. 

Although I found most of the material presented 
in the species accounts to be accurate, I ran across 
several reported “facts” that in my opinion. 


478 


December 2005 


Book Review 


479 


amounted to unsubstantiated and rather far-reach- 
ing speculations. Some examples include a state- 
ment that evidence supports that mated pairs of 
Rough-legged Hawks {Buteo lagopus) perch togeth- 
er on their wintering grounds; statements or im- 
plications that several species of raptors covered in 
the volume relatively commonly exhibit coopera- 
tive hunting; that Gyrfalcons {Falco rusticolus) are 
faster than Peregrine Falcons {F. peregrinus); and 
that the food habits of several small raptors include 
large birds and mammals, such as waterfowl (An- 
atidae), grouse (Tetraoninae) , raccoons {Procyon lo- 
tor), woodchucks (Marntota monax), and hares (Le- 
pus spp.). The latter is true, although most of these 
relatively large preys are taken rarely and most like- 
ly involve very-young animals or carrion. This clar- 
ification is generally not included and the way ac- 
counts are worded, the text implies that such large 
prey are just as commonly taken as small rodents. 
The layout of the book, with citations lumped by 
chapter, makes tracking down the specific sources 
for these far-reaching speculations and implica- 
tions nearly impossible. As a scientist, I found this 
aggravating. 


Besides my mild complaints indicated above, the 
species accounts were concise and informative. 
Within each account, the known longevity record 
for each species is reported, which I found to rep- 
resent an interesting anecdote. Each account is il- 
lustrated with one or more high-quality black and 
white photos of the featured species. For the most 
part, the volume is well-edited and I found only a 
few typographical errors scattered about the text. 

Generally, I feel this volume would make an ex- 
cellent primer for an amateur or beginning stu- 
dent interested in North American birds of prey. 
The presentation of the basic natural history and 
promotion of recreational hawk watching to the 
beginning student of birds is clearly the intended 
target of this publication. I would recommend tbis 
volume to a high school or underclass university 
student that expresses an interest in raptors. Also, 
this book would make an excellent resource for 
local public libraries throughout eastern North 
America. — James C. Bednarz, Department of Bio- 
logical Sciences, Arkansas State University, P.O. 
Box 599, Jonesboro, AR 72467 U.S.A. 


J Raptor Res. 39(4);480-483 
© 2005 The Raptor Research Foundation, Inc. 


Journal of Raptor Research 

INFORMATION FOR CONTRIBUTORS 


The Journal of Raptor Research (JRR) publishes 
original research reports and review articles about 
the biology of diurnal and nocturnal birds of prey. 
All submissions must be in English, but contribu- 
tions from anywhere in the world are welcome. 
Manuscripts are considered with the understand- 
ing that they have not been published, submitted 
or accepted for publication elsewhere. Manuscripts 
are subjected to peer review for evaluation of their 
significance and soundness, and edited to improve 
communication between authors and readers. De- 
cisions of the editor are final. 

Material is published as feature articles, short 
communications (usually not longer than four 
printed pages), and letters (see recent issue of the 
JRR for examples) . Submissions that adhere closely 
to the JRR’s format greatly enhance the efficiency 
and cost of the editorial and publishing processes. 
Author’s efforts in this regard are deeply appreci- 
ated by the editorial staff. 

When submitting scholarly papers, send the orig- 
inal and three copies, a completed checklist (see 
below), and a cover letter that includes: (1) a state- 
ment that the data in the manuscript have not 
been published or accepted for publication in the 
same form, and have not been submitted simulta- 
neously elsewhere, (2) the name and address of 
the corresponding author (in multiauthored pa- 
pers) including any temporary addresses where 
that author will be during the review process (also 
the phone number and, if possible, a FAX number 
and e-mail address of the corresponding author), 
and (3) if applicable, any special instructions. Au- 
thors may also suggest potential reviewers. 

If the manuscript submitted was produced on a 
word processor, also send a diskette (3 1/2") or CD 
containing a single file that is identical with the 
printed copy. The electronic copy should be sup- 
plied as an IBM-compatible text file (Word or 
WordPerfect). Optional electronic submissions are 
encouraged (see General Instructions below). 

Manuscripts are accepted upon the condition 
that the revision must be returned to the editor 
within 60 days. Manuscripts held longer will lose 


their priority and may be treated as new submis- 
sions. The editor should be notified if extenuating 
circumstances prevent a timely return of the man- 
uscript. 

Authors will receive proofs of their articles prior 
to publication. Proofs must be read carefully to 
correct any printer errors and returned by the fast- 
est mail within two days of receipt TO THE EDI- 
TOR. Changes in typeset text are expensive and 
authors making changes, not due to printer error, 
will be billed for the costs ($3.50 US per change). 
A reprint order will accompany page proofs to en- 
able authors to buy reprints. Costs of reprints are 
the author’s responsibility and payment for re- 
prints ordered must accompany the order form. 
Both must be sent TO THE EDITOR. 

Publication is expensive and member dues do 
not cover the entire cost of producing the JRR. 
Hence, the Raptor Research Foundation, Inc. ex- 
pects that authors defray the high costs of publi- 
cation through payment of page costs (currently 
$115.00 U.S. per page). Authors who are not as- 
sociated with a research institution or simply do 
not have access to such grants may request the 
page charges be waived. Such a request can only 
be approved if the author is a member of the Foun- 
dation and the article is short. Payments of 
amounts less than the full page charges will be ac- 
cepted. Authors of long manuscripts are expected 
to pay publishing costs. It is unlikely that articles 
longer than 10 printed pages or 18 typewritten 
pages including tables and illustrations can be pub- 
lished without full payment. Authors employed hy 
government agencies, universities, or firms that 
will meet reprint and page charges may forward a 
statement to the editor indicating intent to pay. 
Upon receipt of such a statement, reprints will be 
mailed to the author and the agency will be billed 
with the understanding that payment will be made 
within 30 days. All checks should be made payable 
to the Raptor Research Foundation, Inc. All per- 
sonal payments toward publication costs are tax- 
deductible in the United States. 


480 


December 2005 


Information for Contributors 


481 


Journal of Raptor Research 

CHECKLIST FOR PREPARATION OF MANUSCRIPTS 
{check items and submit with manuscript) 


I. General Instructions 

(Consult recent issues for additional guidance on format) 

I I Type manuscripts on one side of either 216 X 2*78 
mm (8.5 X 11") or standard international size (210 
X 297 mm) good quality paper (do not use erasable 
or lightweight paper) . Word-processor-generated 
manuscripts must be done with a letter-quality or near- 
letter-quality printer. DOUBLE SPACE THROUGH- 
OUT including title, text, tables, figure legends, and 
literature cited. 

D Electronic submission. Submit e-mail message with 
two attached file copies of manuscript: (1) pdf file 
(must include figures) and (2) word processing file 
(Word Perfect or MS Word files accepted). Follow 
same format guidelines as for standard mail submis- 
sion. Letter of transmittal should be included in the 
e-mail message. E-mail to: journalofraptorresearch® 
juno.com 

□ Give the scientific name at the first mention of a spe- 
cies, both in the abstract and in the article. Scientific 
names of birds should follow the usage of the AOU 
Check-list of North American Birds (7th. Ed. 1998 and 
subsequent supplements in the Auk) or an authorita- 
tive source corresponding to other geographic re- 
gions. Do not give subspecific identification unless it 
is pertinent. Capitalize first letter of words in com- 
plete common names for birds. Use lower case for all 
other common names. 

□ Use American spelling and Webster’s Tenth New Colle- 
giate Dictionary (1996, Merriam-Webster, Inc.) as a 
spelling authority. 

□ Leave at least a 25 mm (1") margin on all sides. Avoid 
hyphens or dashes at ends of lines; do not divide a 
word at the end of a line. 

n Use a nonproportional font of at least elite size (4.7 
characters/cm = 12 characters/inch) or 12 point, 
preferably Courier. DO NOT USE RIGHT JUSTIFI- 
CATION-LEAVE RIGHT MARGIN RAGGED. 

□ Use italic type for addresses, scientific names, journal 
names, and third level headings. 

□ Type last name(s) of author (s) and page number in 
upper right-hand corner of page 2 and all following 
pages. 

□ Cite each figure and table in the text. Do not repeat 
material in two forms (i.e., in text and table, or table 
and figure). Organize text, as far as possible, so that 
tables and figures are cited in numerical order. 

□ Use “Figure” only to start a sentence; otherwise 
“Fig.” if singular, “Figs.” if plural (e.g.. Fig. 1; Figs. 2, 
3; Figs. 4-6). 

□ Use metric units throughout. 

□ Use these abbreviations without spelling out: hr, min, 
sec, yr, mo, wk, d, km, cm, mm; designate temperature 
as 32°C. 


□ Use “continental” dating (e.g., lOJuly 1993, 1-3 June, 
1 1 May to 1 1 June) . 

□ Use 24-hour clock (e.g., 0800 H, 1345-1400 H) 

D Write out numbers one to nine unless a measurement 
(e.g., four birds, 3 km, 40 sites, 6 yr). Use 1000 and 
10 000; 0.15 instead of .15; % instead of percent. 

G Each reference cited in text must be listed in the Lit- 
erature Cited section, and vice versa. Double check 
the accuracy of all entries— THE EDITORIAL STAFF 
CANNOT DO THIS FOR YOU. 

□ Literature citations in the text are as follows: 

a. One authorG^ones (1993) or (Jones 1993) 

b. Two authons-Smith and Jones (1991) or (Smith 
and Jones 1991) 

c. Three or more authors-Hernandez et al. (1990) or 
(Hernandez et al. 1990) 

d. Manuscripts accepted for publication but not yet 
published-Howard (in press) or (Howard in press) 

e. Unpublished materials-K. Jacobson (unpubl 
data); (K. Jacobson pers. comm.); or K. Jacobson 
(pers. comm.) -do not place in the Literature Cited 
section. 

f. When citing several references within parentheses, 
separate with commas and put in chronological or- 
der, oldest first). 

g. For manuscripts submitted as letters, place cita- 
tions in text in abbreviated form, e.g., (LG. Birds 
1993, /. Raptcyr Res. 27:45-50) . 

D Assemble manuscripts for regular articles in this or- 
der: (1) title page, (2) abstract page, (3) text, (4) ta- 
bles, (5) figure legends, (6) figures. DO NOT STA- 
PLE. 

□ Avoid any unnecessary or special formatting. 

II. Title Page 

□ Place full title 6-8 lines below top of page in all 
capital letters. Below title, center author’s 
name(s) in all capital letters and address (es) 
followed by a running title (short title) not to 
exceed 30 characters. If the author(s) is/are 
currently at another location from where the 
work was done, use superscript number (s) fol- 
lowing author (s) name(s) to indicate current 
address in footnote at bottom of the page. In 
multiauthored papers, indicate the author re- 
sponsible for correspondence and requests for 
reprints. Give phone number and, if possible, 
FAX number and e-mail address of the corre- 
sponding author. 

III. Abstract/summary 

□ For regular articles, include an abstract of about 
250 words in one paragraph that is completely 


482 


Information for Contributors 


VoL. 39, No. 4 


without reference to the text. Be concise, in- 
clude the paper’s purpose, but emphasize the 
results. Statements like “results will be dis- 
cussed” are not appropriate. The abstract will 
also be published in Spanish. Authors fluent in 
both languages are encouraged to include both 
versions, otherwise the JRR will provide the 
Spanish translation. 

□ Include five to seven key words for indexing af- 
ter the abstract. 

□ Short communications will be printed with a 
Spanish summary only. Authors must provide 
an English summary to be translated into Span- 
ish unless they are fluent in Spanish. 

□ Avoid citing references in the abstract. If they 
must be cited, include journal name, volume, 
pages, and year, all in parentheses, 

IV. Text 

□ Follow instructions in section I. 

n Main headings are all capital letters and flush 
with left margin. 

□ Typical main headings for regular articles are: 
METHODS, RESULTS, and DISCUSSION. An 
introduction begins the text but does not have 
a heading. 

□ Put second-level headings in bold. Use normal 
indentation and capitalize first letter of each 
word in the second-level headline except prep- 
ositions and articles. 

□ Put third-level headings in italics. Capitalize first 
letter of first word only. 

□ Short communications and letters may or may 
not have headings within the text depending 
upon the need. 

V. LiTEaiATURE Cited 

n Type references in large and small capital let- 
ters, including all authors’ names (e.g., Ne- 
meth, N.M. AND J.L. Morrison. 2002. Natal dis- 
persal of the Crested Caracara . . . etc. . . .). 

□ Do not put space between initials. 

n Verify all entries against original sources includ- 
ing diacritical marks and spelling in languages 
other than English. Capitalize all nouns in Ger- 
man. 

□ Cite references in alphabetical order by hrst au- 
thor’s surname. References by a single author 
precede multiauthored works by the same se- 
nior author regardless of date. 

□ List works by the same author (s) chronologi- 
cally, beginning with the oldest. 


□ Use six hyphens when the author is the same 
as in the preceding citation. 

□ On-line sources should be cited if necessary and 
listed in the Literature Cited. Include au- 
thor (s), year, appropriate “title” of web site, 
publisher or sponsor of web site, web site ad- 
dress, and date last accessed in parentheses. 
Example: 

Sauer, J.R., J.E. Hines, and J. Fallon. 2005. 
The North American breeding bird survey re- 
sults and analysis, 1966-2004. Version 2005.2 
USGS Patuxent Wildlife Research Center, 
Laurel, MD U.S.A. http://www.mbr-pwrc. 
usgs.gov/bbs/bbs.html (last accessed 31 De- 
cember 2005). 

□ “In press” citations must have been accepted 
for publication and must include date, volume 
number, and the name of the journal or pub- 
lisher. 

□ Initials of second, third, and . . . authors pre- 
cede their surname. 

□ Abbreviate journal names according to the Se- 
rial Sources for the BIOSIS Data Base (published 
annually by the BioSciences Information Ser- 
vice). 

O Do not list personal communications and un- 
published reports. 

VI. Tables 

(Tables are expensive to print-keep them to a min- 
imum and put each on a separate page-try to de- 
sign them to fit a single column.) 

□ Double space throughout. Assign each table an 
Arabic number followed by a period. 

□ Table headings should be typed in large and 
small capital letters. 

D Use same size of type as in text. 

□ Indicate footnotes by lowercase superscript let- 
ters. 

□ Do not use vertical lines. 

VII. Figure Legends 

□ Print all figure legends on one page, double 
spaced. 

□ Number using Arabic numbers consecutively m 
the same order the figures appear in the text 
(i.e., Figure L, Figure 2., etc.). 

VIII. Preparation of Illustrations 

(Illustrations are referred to as figures and include 

drawings, graphs, and black and white half-tones 


December 2005 


Information for Contributors 


483 


[photographs] . CONSULT THE EDITOR IN AD- 
VANCE ABOUT COLOR.) 

D Use professional standards in preparing figures; 
their reproduction in the JRR is virtually iden- 
tical to what is submitted. Consult issues of JRR 
for examples and Steps Toward Better Scientific Il- 
lustrations (Allen Press, P.O. Box 368, Lawrence, 
KS 66044) for more information. 

n Plan figures to fit proportions in the JRR, pref- 
erably for a single column-printed size is 72 
mm for single column width, 148 mm for full 
page width or 195 mm for lengthwise figures. 
Figures should be submitted no smaller than 
the final size nor larger than twice the final size 
(on paper no larger than 216 X 278 mm (8.5 
X 11") or standard international (210 X 297 
mm). 

□ All graphics and images should be scanned at a 
minimum resolution of 300 pixels per inch 
(ppi). Line art should be scanned at 1200 ppi. 
Low resolution figures or graphics are not ac- 
ceptable. 

□ Figure text must be a plain (sans serif) typeface 
(e.g., Helvetica), not compressed, and large 
enough so that it will be as large as the text type 
(8—10 point) when in print. 

□ Photographs must be sharp, high-contrast, 


glossy prints approximately the size that they 
will appear in print. If several photographs are 
to be included in one figure, group them butt- 
ed together with no space between. 

□ Use the same style of lettering and presentation 
for all figures. Capitalize each word of axes titles 
except prepositions and articles. 

IX. What to Send 

□ Cover letter. 

□ Copy of this checklist completed. 

□ Original and three copies of manuscript and il- 
lustrations. 

□ Diskette containing a text file of the manuscript 
text, tables, and figures (file in MS Word or 
WordPerfect) . 

□ Electronic submissions are encouraged. See sec- 
tion I. 

□ Submit to: 

Cheryl R. Dykstra, Editor 
Raptor Environmental 
7280 Susan Springs Drive 
West Chester, OH 45069-3696 U.S.A. 

More information: 

Telephone: (513) 779-1744 

E-mail: journalofraptorresearch@juno.com 


J Raptor Res. 39(4):484-491 
© 2005 The Raptor Research Foundation, Inc. 


Index to Volume 39 

Prepared By Autumn A. Farless 


The index includes references to general topics, common names, keywords, and authors. Reference is also made 
to book reviews and letters. Taxa other than raptors are included where referenced by authors. 


A 

Abundance, 80-83 
relative, 466-471 
Accipiter cooperii, 109 

gentilis, 210-221, 222-228, 229-236, 237-246, 247-252, 
264-273, 274-285, 286-295, 296-302, 303-309, 
310-323, 32^334, 335-341, 342-350, 439-444 
atricapillus, 192-209 
laingi, 253-263 
Activity centers, 253-263 
Adaptive kernel, 253-263 
Aegolius acadicus brooksi, 134-141 

Agostini, Nicolantonio, Are earlier estimates of Accipitri- 
formes crossing the channel of Sicily (Central Med- 
iterranean) during spring migration accurate?, 184— 
186 

Agricultural areas, 55-60 
Agriculture, 429-438 
Alarm call, 475-476 

Alencar Carvalho, Carlos Eduardo, see Mendes de Car- 
valho Filho, Eduardo Pio 
Allen, Deborah J., see DeCandido, Robert 
Anchor-bolt ladder, 109 

Andersen, David E., Stephen DeStefano, Michael I. Gold- 
stein, Kimberly Titus, Cole Crocker-Bedford, John J. 
Keane, Robert G. Anthony, Robert N. Rosenfield, 
Technical review of the status of Northern Goshawks 
in the western United States, 192-209 
Andersen, David E., see Boal, Clint W. 

Andersen, David E,, see Smithers, Brett L. 

Andrade, Analia, see Sauthier, Daniel E. Udrizar 
Anthony, Robert G., see Andersen, David E. 
Anthropogenic disturbance, 97-101 
Antolin, Michael R, see Bayard de Volo, Shelley 
Aquila clanga, 462-466 
Argentina, 65-69 

Argiielles, Cerafina, see Castellanos, Aradit 
Arizona, 274-285 
Arkansas, 74-79 
Asio otus, 445-453 
Athene cuniculania, 429—438 
noctua, 156-159, 454—457 
Attalea maripa 

B 

Baja California, 472-475 

Baldassarre, Guy A., see Jensen, Wendy J. 


Bayard de Volo, Shelley, Richard T. Reynolds, J. Rick To- 
pinka, Bernie May, and Michael F. Antolin, Popula- 
tion genetics and genotyping for mark-recapture 
studies of Northern Goshawks {Accipiter gentilis) on 
the Kaibab Plateau, Arizona, 286—295 
Bechard, Marc J., see Fairhurst, Graham D. 

Bednarz, J.G., sec Radley, Paul M. 

Bednarz, J.G., A Review of Hawks and Owls of Eastern 
North America, by Donald S. Heintzelman, 2004, 
478-479 

Beier, Paul, see Boyce, Douglas A., Jr. 

Begall, Sabine, The relationship of foraging habitat to 
the diet of Barn Owls ( Tyto alba) from central Ghile, 
97-101 

Belthoff, James R., see Moulton, Colleen E. 

Bennett, Jason R. and P.H. Bloom, Home range and hab- 
itat use by Great Horned Owls {Bubo virginianus) m 
southern California, 119-126 
Bildstein, Keith L., see Jensen, Wendy J. 

Bird, David M., see Laing, Dawn K. 

Bird, David M., see Chubbs, Tony E. 

Black-Hawk, Common, 351-364 
Cuban, 351-364 

Bloom, Peter H., see Bennett, Jason R. 

Bloszyk, Jerzy, see Gwiazdowicz, Dariusz J. 

Boal, Clint W., Preface: Proceedings of the international 
symposium on the ecology and management of 
Northern Goshawks, 189 

Boal, Clint W., Productivity and mortality of Northern 
Goshawks in Minnesota, 222-228 
Boal, Clint W., see Smithers, Brett L. 

Boavista, 80-83 
Bonin Islands, 173-179 
Bootstrapping, 253-263, 274-285 
Botella, Francisco, see Navarro, Joan 
Boyce, Douglas A., Jr., Patricia L. Kennedy, Paul Beier, 
Michael F. Ingraldi, Susie R. MaeVean, Melissa S. Sid- 
ers, John R. Squires, and Brian Woodbridge, When 
are goshawks not there? Is a single visit enough to 
infer absence at occupied nest areas?, 296-302 
Brady, Ryan S., see Moulton, Colleen E. 

Brazil, southeastern, 89-92 
Breeding, 222-228, 229-236 
chronology, 74-79 
cooperative, 92-94 
range, 70-74 


484 


December 2005 


Index to Volume 39 


485 


British Columbia, 1-10, 335-341 
Bubo magellanicus, 163-166 
virginianus, 111-118, 119-126 
Bustamante, Javier, see Rodriguez, Carlos 
Buteo buteo, 466-474 
jamaicensis, 108, 439-444 
platypterus brunnescens, 404—416 
toyoshimai, 173-179 
Buteogallus anthracinus, 351-364 
gundlachii, 351—364 
subtilis, 351-364 

Buzzard, Common, 173-179, 466-474 

C 

California, southern, 119-126 
Calvo, Jose F., see Martinez, Jose E. 

Camera, remote, 303-309 
Camiha, Alvaro, see Garrido, Jose Rafael 
Canary Islands, 186-187 
Caprimulgus macrurus, 106-107 
Capture-recapture, 286-295 
Caracara, Yellow-headed, 94—97 
Caribbean, 94—97 

Carrete, Martina, see Navarro, Joan 

Casado, Eva and Miguel Ferrer, Analysis of reservoir se- 
lection by wintering Ospreys {Pandion haliaetus hal- 
iaetus) in Andalusia, Spain: a potential tool for re- 
introduction, 168-173 

Castellanos, Aradit, Cerafina Arguelles, Federico Salinas- 
Zavala, Alfredo Ortega-Rubio, New nesting record 
and observations of breeding Peregrine Falcons in 
Baja California Sur, Mexico, 472-475 
Cebus olivaceus, 458—461 
Cervera, Francisco, see Garcia, Ana Maria 
Chambers, Carol L., see Gatto, Angela E. 

Chile, 55-60, 97-101 
Chiroptera, 445-453 

Chubbs, Tony E., Matthew J. Solensky, Dawn K. Laing, 
David M. Bird, and Geoff Goodyear, Using a porta- 
ble anchor-bolt ladder to access rock-nesting Osprey, 
103-106 

see Laing, Dawn K. 

Circus spilonotus, 106-107 
CITES, 386-393 
Colorado, 166—168 
Community, 156-159 
Competition, 156-159, 439-444 
Connecticut, 342-350 

Corales Stappung, Ema Soraya, see Rojas, Ricardo A. Fi- 
gueroa 

Crete, 179-183 

Crocker-Bedford, Cole, see Andersen, David E. 

Crozier, Michelle L., The effect of broadcasting Great 
Horned Owl vocalizations on Spotted Owl vocal re- 
sponsiveness, 111-118 


D 

DeCandido, Robert and Deborah J. Allen, First nesting 
of Cooper’s Hawks {Accipiter cooperii) in New York 
City since 1955, 109 

Dekker, Dick and Robert Taylor, A change in foraging 
success and cooperative hunting by a breeding pair 
of Peregrine Falcons and their fledglings, 394-403 
de la Rocha, J.L. Paz, see Garrido, Jose Rafael 
DeLong, John R, Timothy D. Meehan, and Ruth B. 
Smith, Investigating fall movements of hatch-year 
Flammulated Owls {Otus flammeolus) in central New 
Mexico using stable hydrogen isotopes, 19-25 
Departure, 462-466 

Desimone, Steven M. and Stephen DeStefano, Temporal 
patterns of Northern Goshawk nest area occupancy 
and habitat: a retrospective analysis, 310-323 
DeStefano, Stephen, see Andersen, David E. 

DeStefano, Stephen, see Rogers, Andi S. 

DeStefano, Stephen, see Desimone, Steven M. 

DeStefano, Stephen, A review of the status and distribu- 
tion of Northern Goshawks in New England, 342- 
350 

Detectability, 274-285 
Detection rates, 296-302 

Diet, 55-60, 65-69, 80-83, 97-101, 163-166, 173-179, 
179-183, 237-246, 264-273, 303-309, 439-444, 445- 
453 

Digital image analysis, 127-133 
Dispersal, 11-18 
natal, 253-263 
Distance sampling, 237-246 
Distribution, 342-350 
Doyle, Donald D., see McClaren, Erica L. 

Doyle, Frank I., see Mahon, Todd 

E 

Eastham, Chris P. and Mike K. Nicholls. Morphometric 
analysis of large Falco species and their hybrids with 
implications for conservation, 386-393 
Eagle, Bald, 1-10, 11-18 
Booted, 92-94, 159-163 
Greater Spotted, 462-466 
Ecology, 351-364 

Egea, Maria, see Garrido, Jose Rafael 
Eggs, 89-92 

Egyptian Vulture, 186-187 
Elanoides forficatus, 94—97 
Elanus leucurus, 378—385 
Elliott, John E., see Elliott, Kyle Hamish 
Elliott, Kyle Hamish, Christopher E. Gill, and John E 
Elliott, The influence of tide and weather on provi- 
sioning rates of chick-rearing Bald Eagles in Vancou- 
ver Island, British Columbia, 1-10 
Endangered, 404—416 

Enderson, James H., Changes in site occupancy and nest- 


486 


Index to Volume 39 


VoL. 39, No. 4 


ing performance of Peregrine Falcons in Colorado, 
1963-2004, 166-168 
Europe, eastern, 36-54 
Experiment, cross-over, 111-118 

F 

Eairhurst, Graham D. and Marc J. Bechard, Relationships 
between winter and spring weather and Northern 
Goshawk (Accipiter gentilis) reproduction in northern 
Nevada, 229-236 
Falco cherrug, 386—393 
femoralis, 55-60 
naumanni, 12V-133 
nexvtoni, 149-155 
novaezeelandiae, 386-393 

peregrinus, 166-168, 386—393, 394^403, 472—475 
rusticolus, 386-393 
sparverius, 84—88, 94—97, 378—385 
tinnunculus alexandri, 80—83 
Falcon, Aplomado 55-60 
New Zealand, 386-393 

Peregrine, 166-168, 386-393, 394-403, 472-475 
Saker, 386-393 
Falcons, 378-385 
Family break up, 462-466 
Feather, 84-88 
Feeding ecology, 97-101 
Ferland, Cheron, see Forsman, Eric D. 

Ferrer, Miguel, see Casado, Eva 
Fidelity, site, 134—141 
Fish production, 168-173 
Fitness, 210-221 

Fitzsimons, James A., Attempted predation on a Large- 
tailed Nightjar (Caprimulgus macrurus) by an Eastern 
Marsh-Harrier ( Circus spilonotus) in coastal Vietnam, 
106-107 

Fledging-dependency period, 253-263 
success, 26-35 

Food habits, 429-438, 439-444 
niche, 429-438 
breadth, 439-444 
partitioning, 159-163 
Foraging, 439-444 
association, 458-461 
Forest, 159-163 
karst, 404—416 

management, 296-302, 324—334 
Forsman, Eric D., TimmothyJ. Kaminski, Jeffery C. Lewis, 
Kevin J. Maurice, Stan G. Sovern, Cheron Ferland, 
and Elizabeth M, Glenn, Home range and habitat 
use of Northern Spotted Owls on the Olympic Pen- 
insula, Washington, 365-377 
Fortabat, Sofia Heinonen, see Pardihas, Ulyses FJ. 
French, John B., Jr., see Quinn, Michael J.,Jn 
Frye, Graham G., A previously undescribed vocalization 
of the Northern Pygmy-Owl, 476-477 


G 

Gaibani, Giorgia, see Pandolfi, Massimo 
Gangoso, Laura and Cesanjavier Palacios, Ground nest- 
ing by Egyptian Vultures {Neophron perenopterus) in 
the Canary Islands, 186-187 

Garcia, Ana Maria, Francisco Cervera, and Alejandro 
Rodriguez, Bat predation by Long-eared Owls m 
Mediterranean and temperate regions of southern 
Europe, 445-453 
Garcia, David, see Navarro, Joan 

Garrido, Jose Rafael, Alvaro Camina, Mariangela Guinda, 
Maria Egea, Nourdine Mouati, Alfonso Godino, and 
J.L. Paz de la Rocha, Absence of the Eurasian Grif- 
fon {Gyps fulvus) in northern Morocco, 70-74 
Garrido, Orlando H., see Wiley, James W. 

Gatto, Angela E., Teryl G. Grubb, and Carol L. Cham- 
bers, Red-tailed Hawk dietary overlap with Northern 
Goshawks on the Kaibab Plateau, Arizona, 439-444 
Gender determination, 127-133 
Genetic structure, 192-209 
variability, 142-148 
Geographical-catchment area, 19-25 
Gill, Christopher E., see Elliott, Kyle Hamish 
Gimenez, Andres, see Navarro, Joan 
Glaucidium gnoma, 476-477 
Glenn, Elizabeth M., see Forsman, Eric D. 

Godino, Alfonso, see Garrido, Jose Rafael 
Goldstein, Michael L, see Andersen, David E. 

Gonzalez, Carlos, see Martinez, Jose E. 

Gonzalez-Bravo, Betzabeth, see Meraz, Juan 
Goodyear, Geoff, see Chubbs, Tony E. 

Goshawk, Northern, 192-209, 210-221, 222-228, 229- 
236, 237-246, 247-252, 253-263, 264-273, 274-285, 
286-295, 296-302, 303-309, 310-323, 324-334, 335- 
341, 342-350, 439-444 
Gregory, Mark S., see Jensen, Wendy J. 

Griffon, Eurasian, 70-74, 179-183 

Ground nesting, 186-187 

Grubb, Teryl G., see Gatto, Angela E. 

Guinda, Mariangela, see Garrido, Jose Rafael 
Gutierrez, R.J., see Crozier, Michelle L. 

Gwiazdowicz, Dariusz J., Jerzy Bloszyk, Tadeusz Mizera, 
and Piotr Tryjanowski, Mesostigmatic mites (Acari 
Mesostigmata) in White-tailed Sea Eagle nests {Hal- 
iaeetus albicilla) , 60—65 
(dyps fulvus, 70-74, 179-183 
Gyrfalcon, 386—393 

H 

Habitat, 310-323 
modek 404—416 
relations, 192-"209 
use, 119-126, 365-377 
Habits, food, 149^155 
Haliaeetus dlMcilla, 60—65 
leucOcephalus, 1—10, 11-18 


December 2005 


Index to Volume 39 


487 


Harrier, Eastern Marsh, 106-107 
Hatching success, 26-35 
Hawk, Broad-winged, 404-416 
Cooper’s, 109 
Red-tailed, 108, 439-444 
Hawks, 378-385 

Hegyi, Zoltan, see Sasvari, Lajos 

Hengstenberg, Derek W. and Francisco Vilella, Nesting 
ecology and behavior of Broad-winged Hawks in 
moist karst forests of Puerto Rico, 404-416 
Heterogeneity, individual, 210-221 
Hieraaetus pennatus, 92-94, 159-163 
Hoffmann, Gyula, see Matics, Robert 
Holschuh, Carmen I. and Ken A. Otter, Using vocal in- 
dividuality to monitor Queen Charlotte Saw-whet 
Owls (Aegolius acadicus brooksi), 134—141 
Home range, 119-126, 365-377 
Horvath, Gyozo, see Matics, Robert 
Honey-buzzard, European, 184—186 
Hunting, adult, 394-403 
cooperative, 394—403 
fledgling, 394-403 
tandem, 394—403 
Hybrids, falcon, 386-393 

I 

IberVNeembucu wetlands, 65-69 
Ictinia mississippiensis, 108 
Idaho, 429-438 

Immature movements, 253—263 
Individual identification, 286—295 

Ingraldi, Michael F., A skewed sex ratio in Northern Gos- 
hawks: is it a sign of a stressed population?, 247-252 
Ingraldi, Michael F., see Boyce, Douglas A., Jr. 

Ingraldi, Michael F., see Rogers, Andi S. 

Introduced animals, 173—179 
Introgression, 142-148 
Iverson, W.F., 102-103 

J 

Juvenile, 11-18 

Jensen, Wendy J., Mark S. Gregory, Guy A. Baldassarre, 
Francisco Vilella, and Keith L. Bildstein, Raptor 
abundance and distribution on the Llanos wedands 
of Venezuela, 417-428 
Joy, Suzanne M., see Reynolds, Richard T. 

K 

Kaibab Plateau, 210-221 

Kaminski, TimmothyJ., see Forsman, Eric D. 

Kato, Yuka and Tadashi Suzuki, Introduced animals in 
the diets of the Ogasawara Buzzard, an endemic in- 
sular raptor in the Pacific Ocean, 173-179 
Keane, John J., see Andersen, David E. 

Kennedy, Patricia L., see Boal, Clint W. 

Kennedy, Patricia L., see McClaren, Erica L. 


Kennedy, Patricia L., see Boyce, Douglas A., Jr. 

Kestrel, Alexander’s, 80-83 
American, 84-88, 94-97, 378-385 
Common, 80-83 
Lesser, 127-133 
Madagascar, 149-155 
Kite, Black, 184-186 
Gray-headed, 89-92 
Mississippi, 108 
Swallow-tailed, 94-97 
White-tailed, 378-385 
Klein, Akos, see Matics, Robert 
Kowalski, Jan, see Meyburg, Bernd-U. 

L 

Labrador, 11-18 

Laing, Dawn K., David M. Bird, and Tony E. Chubbs, First 
complete migration cycles for juvenile Bald Eagles 
{Haliaeetus leucocephalus) from Labrador, 11-18 
Laing, Dawn K., see Chubbs, Tony E. 

Landscape change, 310-323 

Leptodon cayanensis, 89-92 

Lewis, Jeffery C., see Forsman, Eric D. 

Livezey, Kent B., Iverson (2004) on Spotted Owls and 
Barred Owls: comments on methods and conclu- 
sion, 102-103 
Llanos, 417-428 

M 

Maciorowski, Grzegorz, see Meyburg, Bernd-U. 

MacVean, Susie R., see Boyce, Douglas A., Jr. 

Mahon, Todd and Frank I. Doyle, Effects of timber har- 
vesting near nest sites on the reproductive success of 
Northern Goshawks {Accipiter gentilis) , 335-341 
Maine, 342-350 

Management, adaptive, 335-341 
Martinez, Jose Antonio, see Zuberogoitia, Inigo 
Martinez, Jose E., Carlos Gonzalez, and Jose F. Calvo, Co- 
operative nesting by a trio of Booted Eagles {Hieraae- 
tus pennatus) , 92-94 

Martinez, Jose Enrique, see Zuberogoitia, Inigo 
Martinez, Jose E. and Jose F. Calvo, Prey partitioning be- 
tween mates in breeding Booted Eagles {Hieraaetus 
pennatus), 159-163 

Martinez-Cruz, Begona, see Rodriguez, Carlos 
Massachusetts, 342-350 

Matics, Robert, Sandor Varga, Balazs Opper, Akos Klein, 
Gydzo Horvath, Alexandre Roulin, Peter Putnoky, 
and Gyula Hoffmann, Partitioning of genetic 
(RAPD) variability among sexes and populations of 
the Barn Owl {Tyto alba) in Europe, 142-148 
Maurice, Kevin J., see Forsman, Eric D. 

Mauritia Jlexuosa, 458-461 
May, Bernie, see Bayard de Volo, Shelley 
McClaren, Erica L., Patricia L. Kennedy, and Donald D 
Doyle, Northern Goshawk {Accipiter gentilis laingi) 


488 


Index to Volume 39 


VoL. 39, No. 4 


post-fledging areas on Vancouver Island, British Co- 
lumbia, 253—263 

McNabb, F.M. Anne, see Quinn, Michael J., Jr. 
Mediterranean basin, 445-453 
central, 184—186 

Meehan, Timothy D., see DeLong, John P. 

Mendes de Carvalho, Gustavo Diniz, see Mendes de Car- 
valho Filho, Eduardo Pio 

Mendes de Carvalho Filho, Eduardo Pio, Gustavo Diniz 
Mendes de Carvalho, and Carlos Eduardo Alencar 
Carvalho, Observations of nesting Gray-headed Kites 
{Leptodon cayanensis) in southeastern Brazil, 89-92 
Meraz, Juan and Betzabeth Gonzalez-Bravo, First summer 
records of Ospreys {Pandion haliaetus) along the 
coast of Oaxaca, Mexico, 18V-188 
Mesostigmata, 60-65 
Mexico, 187-188, 472-475 

Meyburg, Bernd-U., Christiane Meyburg, Tadeusz Mizera, 
Grzegorz Maciorowski, and Jan Kowalski, Family 
break up, departure, and autumn migration in Eu- 
rope of a family of Greater Spotted Eagles {Aquila 
clanga) as reported by satellite telemetry, 462-466 
Meyburg, Christiane, see Meyburg, Bernd-U. 

Migration, 11-18, 94—97, 462—466 
counts, 184—186 
patterns, 19—25 
spring, 184-186 

Miller, Ikarl E., Red-tailed Hawk depredates Mississippi 
Kite nestling at dawn, 108 
Milvago chimachima, 94—97 
Milvus migrans, 184-186 
Minguez, Eduardo, see Navarro, Joan 
Minnesota, 222-228, 264-273 
Mites, 60-65 

Mizera, Tadeusz, see Gwiazdowicz, Dariusz J. 

Mizera, Tadeusz, see Meyburg, Bernd-U. 

Molecular sexing, 286-295 
Molt, 84-88 
preformative, 378-385 
Monitoring, 274-285 
raptor, 324-334 

Monkeys, wedge-capped capuchin, 458-461 
Morocco, 70-74 
Morphometric, 386-393 
Mortality, 222-228 

Mouati, Nourdine, see Garrido, Jose Rafael 
Moulton, Colleen E., Ryan S. Brady, and James R. Belt- 
hoff, A comparison of breeding season food habits 
of Burrowing Owls nesting in agricultural and non- 
agricultural habitat in Idaho, 429-438 

N 

Navarro, Joan, Eduardo Minguez, David Garcia, Carlos 
Villacorta, Francisco Botella, Jose Antonio Sanchez- 
Zapata, Martina Carrete, and Andres Gimenez, Dif- 
ferential effectiveness of playbacks for Little Owls 


{Athene noctua) surveys before and after sunset, 454— 
457 

Negro, Juan Jose, see Rodriguez, Carlos 
Neophron percnopterus, 1 86-1 87 
Neotropics, 417-428 
Nest alternate, 274-285 
area, 296-302, 335-341 
historical, 310-323 
biology 60-65, 89-92 
depredation, 108 
structure, 89-92 
success, 404-416 
Nesting, 109 

biology, 89-92, 149-155 
record, 472-475 

Nesting structures, artificial, 74-79 

Nestling, 89-92, 127-133, 247-252 

Nests, 149-155 

Nevada, 229-236 

New England, 342-350 

New Hampshire, 342-350 

New York City, 109 

Nicholls, Mike K., see Eastham, Chris P. 

Nightjar, Large-tailed, 106-107 

Nijman, Vincent, Tineke G. Prins, andJ.H. (Hans) Reu- 
ter, Timing and abundance of migrant raptors on 
Bonaire, Netherlands Antilles, 94-97 
Noon, Barry R., see Salafsky, Susan R. 

O 

Oaxaca, 187-188 
Occupancy, 296-302 
nest site, 324-334 
rate, 166-168 

Ogasawara Islands, 173-179 
Olympic Peninsula, 365-377 

Ontiveros, Diego, Abundance and diet of Alexander’s 
Kestrel {Falco tinnunculus alexandri) on Boavista Is- 
land (Archipelago of Cape Verde), 80-83 
Opper, Balazs, see Matics, Robert 
Oregon, 310-323 

Ortega-Rubio, Alfredo, see Castellanos, Aradit 
Osprey, 103-106, 168-173, 187-188 
Otter, Ken A., see Holschuh, Carmen I. 

Ottinger, Mary Ann, see Quinn, Michael J., Jr. 

Otus Jlammeolus, 19—25 
scops, 156-159 

Owl, Barn, 65-69, 74-79, 97-101, 142-148, 156-159, 
163-166 

Barred, 102-103 
Burrowing, 429-438 
California Spotted, 111-118 
Flammulated, 19-25 
Great Horned, 111-118, 119-126 
Little, 156-159, 454-457 
Long-eared, 445-453 


December 2005 


Index to Volume 39 


489 


Magellanic Horned, 163-166 
Northern Saw-whet, 134-141 
Northern Spotted, 184—186, 365-377 
Rufous-legged, 163-166 
Scops, 156-159 
Tawny, 26-35, 156-159 

P 

Palacios, Cesar-Javier, see Gangoso, Laura 
Palms, Attalea maripa, 458-461 
Pandion haliaetus, 103-106, 168-173, 187-188 
Pandolfi, Massimo, Alessandro Tanferna, and Giorgia 
Gaibani, Seasonal patterns of Common Buzzard (Bu- 
teo buteo) relative abundance and behavior in Pollino 
National Park, Italy, 466-471 
Paraguay, 65—69 

Pardihas, Ulyses F.J., see Sauthier, Daniel E. Udrizar 
Pardihas, Ulyses F.J., Pablo Teta, and Sofia Heinonen For- 
tabat. Vertebrate prey of the Barn Owl ( Tyto alba) in 
subtropical wetlands of northeastern Argentina and 
eastern Paraguay, 65-69 
Parental age, 26-35 
condition, 26-35 

Patla, Susan M., Monitoring results of Northern Goshawk 
nesting areas in the greater Yellowstone ecosystem: 
is decline in occupancy related to habitat change?, 
324-334 
PGA, 386-393 
Pellet analysis, 179-183 
Pernis apivorus, 184—186 
Playback effectiveness, 454-457 
Plumage, 84-88 
Poland, 60-65 
Polygamy, 92-94 
Polygyny, 92-94 
Population change, 166-168 
status, 36-54 
trend, 192-209, 229-236 
Predation, 156-159 
attempted, 106—107 
Predator-prey dynamics, 237-246 
Predictive model, 168-173 
Prey biomass, 65-69 
delivery, 404—416 
rate, 303-309 
density, 237-246 
diversity, 264-273 
provisioning, 159-163 
Principal component analysis, 386-393 
Pnns, Tineke G., see Nijman, Vincent 
Probability of identity, 286-295 

Productivity, 1-10, 74-79, 149-155, 166-168, 222-228, 
237-246 

Puerto Rico, 404—416 

Putnoky, Peter, see Matics, Robert 

Pygmy-Owl, Northern, 476-477 


Pyle, Peter, First-cycle molts in North American Falconi- 
formes, 378-385 

Q 

Quinn, Micheal Jr., John B. French, Jr., F.M. Anne 
McNabb, and Mary Ann Ottinger, The role of thy- 
roxine on the production of plumage in the Amer- 
ican Kestrel {Falco sparverius) , 84—88 

R 

Rabearivony, Jeanneney, see Rene de Roland, Lily-Arison 
Radiotelemetry, 365-377 

Radley, Paul M. and James C. Bednarz, Artificial nest 
structure use and reproductive success of Barn Owls 
in northeastern Arkansas, 74-79 
RAPD, 142-148 

Razafimanjato, Gilbert, see Rene de Roland, Lily-Arison 
Recruitment, 210-221 
Reintroduction, 168-173 

Rene de Roland, Lily-Arison, Jeanneney Rabearivony, 
Harilalaina Robenarimangason, Gilbert Razafiman- 
jato, and Russell Thorstrom, Breeding biology and 
food habits of the Madagascar Kestrel {Falco newtom) 
in northeastern Madagascar, 149-155 
Reproduction, 274-285 

Reproductive success, 74-79, 102-103, 166-168, 210-221, 
335-341 

Reservoirs, 168-173 
Response, heterospecific, 111-118 
Reuter, J.H. (Hans), see Nijman, Vincent 
Reversed size dimorphism (RSD), 159-163 
Review, 192-209 

Reynolds, Richard T., In raemoriam: Suzanne Merideth 
Joy, 190-191 

Reynolds, Richard T., see Wiens, J. David 
Reynolds, Richard T., see Salafsky, Susan R. 

Reynolds, Richard T, J. David Wiens, Suzanne M. Joy, 
and Susan R. Salafsky, Sampling considerations for 
demographic and habitat studies of Northern Gos- 
hawks, 274—285 

Reynolds, Richard T, see Bayard de Volo, Shelley 
RGB values, 127-133 
Rhode Island, 342-350 

Robenarimangason, Harilalaina, see Rene de Roland, 
Lily-Arison 

Rodents, sigmodontine, 163-166 
Rodriguez, Alejandro, see Garcia, Ana Maria 
Rodriguez, Carlos, Javier Bustamante, Begona Martinez- 
Cruz, and Juan Jose Negro, Evaluation of methods 
for gender determination of Lesser Kestrel nestlings, 
127-133 

Rogers, Andi S., Stephen DeStefano, and Michael F. In- 
graldi. Quantifying Northern Goshawk diets using 
remote cameras and observations from blinds, 303- 
309 

Rojas, Ricardo A. Figueroa and Ema Soraya Corales Stap- 
pung, Seasonal diet of the Aplomado Falcon {Falco 


490 


Index to Volume 39 


VoL. 39, No. 4 


femaralis) in an agricultural area of Araucania, south- 
ern Chile, 55—60 

Rosenfield, Robert N., see Andersen, David E. 

Roulin, Alexandre, see Matics, Robert 

S 

Salafsky, Susan R., Richard T. Reynolds, and Barry R. 
Noon, Patterns of temporal variation on goshawk re- 
production and prey resources, 237-246 
Salafsky, Susan R., see Reynolds, Richard T. 
Sahnas-Zavala, Federico, see Castellanos, Aradit 
Sampling, 274-285 
repeated, 296-302 

Sanchez-Zapata, Jose Antonio, see Navarro, Joan 
Sarcoramphus papa, 458-461 

Sasvari, Lajos and Zoltan Hegyi, Effects of breeding ex- 
perience on nest-site choice and the reproductive 
performance of Tawny Owls {Strix aluco), 26-35 
Satellite telemetry, 11-18, 462-466 

Sauthier, Daniel E. Udrizar, Analia Andrade, and Ulyses 
FJ. Pardihas, Predation of small mammals by Rufous- 
legged Owl, Barn Owl, and Magellanic Horned Owl 
in Argentinian Patagonia Forests, 163-166 
Savanna, 417-428 

Schlee, Marsha A., King Vultures {Sarcoramphus papa) for- 
age in moriche and cucurit palm stands, 458-461 
Sea Eagle, White-tailed, 60-65 
Seamons, Mark E., see Crozier, Michelle L. 

Sex allocation, 247-252 
ratio, 247-252 
Sicily, Channel of, 184—186 
Siders, Melissa S., see Boyce, Douglas A., Jr. 

Smith, Ruth B., see DeLong, John P. 

Smithers, Brett L., Clint W. Boal, and David E. Andersen, 
Northern Goshawk diet in Minnesota: an analysis us- 
ing video recording systems, 264—273 
Solensky, Matthew J., see Chubbs, Tony E. 

Severn, Stan G., see Forsman, Eric D. 

Spain, southern, 168-173 

Squires, John R., see Boyce, Douglas A., Jr. 

Squirrel, red, 264-273 
Stable-hydrogen isotopes, 19-25 
Status, 192-209, 342-350 
Stopover, 11-18 
Stnx aluco, 26-35, 156-159 
occidentalis caurina, 365-377 
occidentalis, 102-103, 111-118 
rufipes, 163-166 
varia, 102—103 
Subspecies, 142-148 
Insular 173—179 
Success, reproductive, 166—168 
Survey, auditory, 111-118 
method, 454—457 
population, 454—457 
roadside, 417-428, 466-471 


techniques, 324—334 
Suzuki, Tadashi, see Kato, Yuka 

T 

Tagging, genetic, 286-295 

Tamiasciurus hudsonicus, 264—273 

Tanferna, Alessandro, see Pandolfi, Massimo 

Taxonomy, 351-364 

Taylor, Robert, see Dekker, Dick 

Techniques, 103-106 

Territoriality, 111-118 

Territory occupancy, 274—285 

Teta, Pablo, see Pardihas, Ulyses FJ. 

Thorstrom, Russell, see Rene de Roland, Lily-Arison 
Threats, 36-54 
Thyroxine, 84—88 
Tides, 1-10 

Timber harvesting, 335-341 

Titus, Kimberly, see Andersen, David E. 

Topinka, J. Rick, see Bayard de Volo, Shelley 
Trio, 92-94 

Trophic plasticity, 445-453 

Tryjanowski, Piotr, see Gwiazdowicz, Dariusz J. 

Turan, Levent, The status of diurnal birds of prey in Tur- 
key, 36—54 
Turkey, 36-54 

Tyto alba, 65-69, 97-101, 142-148, 156-159, 163-166 
pratincola, 74-79 

U 

United States, northeastern, 342-350 
western, 192-209 

V 

Varga, Sandor, see Matics, Robert 
Venezuela, 417-428 
Vermont, 342-350 
Video surveillance, 303—309 
Vietnam, 106—107 

Vilella, Francisco, see Jensen, Wendy J. 

Vilella, Francisco, see Hengstenberg, Derek W. 
Villacorta, Carlos, see Navarro, Joan 
Vocal individuality, 134—141 
Vocalization, 476—477 
Vulture, King, 458-461 

W 

Washington, 365-377 
Weather, 1-10, 229-236 
condition, 26-35 
Wetlands, 417-428 

Wiens, J. David and Richard T. Reynolds, Is fledging suc- 
cess a reliable index of fitness in Northern Gos- 
hawks?, 210-221 

Wiens, J. David, see Reynolds, Richard T. 

Wiley, James W. and Orlando H. Garrido, Taxonomic sta- 


December 2005 


Index to Volume 39 


491 


tus and biology of the Cuban Black-Hawk, Buteogallus 
anthracinus gundlachii (Aves: Accipitridae) , 351—364 
Wintering, 70-74 

Woodbridge, Brian, see Boyce, Douglas A., Jr. 

X 

Xirouchakis, Stavros M., The diet of Eurasian Griffons 
{Gyps fulvus) in Crete, 179-183 


Z 

Zabala, Jabi, see Zuberogoitia, Inigo 

Zuberogoitia, Inigo, Jose Antonio Martinez, Jabi Zabala, 
and Jose Enrique Martinez, Interspecific aggression 
and nest-site competition in a European owl com- 
munity, 156-159 


THE JOURNAL OF RAPTOR RESEARCH 

A QUARTERLY PUBLICATION OF THE RAPTOR RESEARCH FOUNDATION, INC. 

(Founded 1966) 



EDITOR IN CHIEF 
James C. Bednarz 


ASSOCIATE EDITORS 


James R. Belthoff 
Clint W. Boal 
Cheryl R. Dykstra 
Michael I. Goldstein 


Joan L. Morrison 
Fabrizio Sergio 
Ian G. Warrentin 
James W. Watson 


BOOK REVIEW EDITOR 

Jeffrey S. Marks 


CONTENTS FOR VOLUME 39, 2005 


Number 1 


The Influence of Tide and Weather on Provisioning Rates of Chick-rearing 
Bald Eagles in Vancouver Island, British Columbia. Kyle Hamish Elliott, 

Christopher E. Gill, and John E. Elliott 1 

First Complete Migration Cycles for Juvenile Bald Eagles {Haliaeetus 

LeUCOCEPHALUS) from Labrador. Dawn K. Laing, David M. Bird, and Tony E. Chubbs ........ 11 

Investigating Fall Movements of Hatch-year Flammulated Owls ( Otus 

FLAMMEOLUS) IN CENTRAL NeW MEXICO USING STABLE HYDROGEN ISOTOPES. 

John P. DeLong, Timothy D. Meehan, and Ruth B. Smith 19 

Effects of Breeding Experience on Nest-site Choice and the Reproductive 
Performance of Tawny Owls (Strix aluco) . Lajos Sasvari and Zoitan Hegyi 26 

The Status of Diurnal Birds of Prey in Turkey. Levent Turan 36 

Short Communications 

Seasonal Diet of the Aplomado Falcon (Falco femoralis) in an Agricultural Area of Araucania, 

Southern Chile. Ricardo A. Figueroa Rojas and Ema Soraya Corales Stappung 55 

Mesostigmatic Mites (Acari: Mesostigmata) in White-tailed Sea Eagle Nests {Haliaeetus albicilla) . 

Dariusz J. Gwiazdowicz, Jerzy Bloszyk, Tadeusz Mizera, and Piotr Tryjanowski 60 

Vertebrate Prey of the Barn Owl ( Tyto alba) in Subtropical Wetlands of Northeastern 

Argentina and Eastern Paraguay. Ulyses FJ. Pardinas, Pablo Teta, and Sofia Heinonen Fortabat 65 

Absence of the Eurasian Griffon {Gyps fulvus) in Northern Morocco. Jose Rafael Garrido, 

Alvaro Camiha, Mariangela Guinda, Marfa Egea, Nourdine Mouati, Alfonso Godino, and 

J.L. Paz de la Rocha 70 

Artificial Nest Structure Use and Reproductive Success of Barn Owls in Northeastern 

Arkansas. Paul M. Radley and James C. Bednarz 74 

Abundance and Diet of Alexander’s Kestrel {Falco tinnunculus alexandri) on Boavista Island 

(Archipelago of Cape Verde) . Diego Ontiveros 80 

The Role of Thyroxine on the Production of Plumage in the American 

Kestrel {Falco sparverius) . Michael J. Quinn, Jr., John B. French, Jr., F.M. Anne McNabb, 

and Mary Ann Ottinger 84 

Observations of Nesting Gray-headed Kites {Leptodon cayanensis) in Southeastern Brazil. 

Eduardo Pio Mendes de Carvalho Filho, Gustavo Diniz Mendes de Carvalho, and Carlos Eduardo 
Aencar Carvalho 89 

Cooperative Nesting by a Trio of Booted Eagles {Hieraaetus pennatus). Jose E. Martmez, 

Carlos Gonzalez, and Jose F. Calvo 92 

Timing and Abundance of Migrant Raptors on Bonaire, Netherlands Antilles. Vincent Nijman, 

Tineke G. Prins, andJ.H. (Hans) Reuter 94 

The Relationship of Foraging Habitat to the Diet of Barn Owls ( Tyto alba) from Central Chile. 

Sabine Begall 97 

Letters 

Iverson (2004) on Spotted Owls and Barred Owls: Comments on Methods and Conclusion. 

Kent B. Livezey 102 


Using a Portable, Anchor-bolt Ladder to Access Rock-nesting Osprey. Tony E. Ghubbs, 

Matthew J. Solensky, Dawn K. Laing, David M. Bird, and Geoff Goodyear 103 

Attempted Predation on a Large-tailed Nightjar ( Caprimulgus macrurus) by an Eastern 

Marsh-Harrier (Circus spilonotus) in Coastal Vietnam. James A. Fitzsimons 106 

Red-tailed Hawk Depredates Mississippi Kite Nestling at Dawn. Karl E. Miller 108 

First Nesting of Cooper’s Hawks (Accipiter cooperu) in New York City Since 1955 . 

Robert DeCandido 1 09 

Manuscript Referees 110 


Number 2 

The Effect of Broadcasting Great Horned Owl Vocalizations on Spotted Owl 

Vocal Responsiveness. Michelle L. Crozier, Mark E. Seamans, and R.J. Gutierrez Ill 

Home Range and Habitat Use by Great Horned Owls {Bubo virginianus) in 

Southern California. Jason R. Bennett and Peter H. Bloom 119 

Evaluation of Methods for Gender Determination of Lesser Kestrel Nestlings. 

Carlos Rodriguez, Javier Bustamante, Begona Martmez-Cruz, and Juan Jose Negro 127 

Using Vogal Individuality to Monitor Queen Charlotte Saw-whet Owls 

(AeGOLIUS ACADICUS BROOKS !) . Carmen I. Holschuh and Ken A. Otter 134 

Partitioning of Genetic (RAPD) Variability among Sexes and Populations 
OF THE Barn Owl ( TyTO alba) in Europe. Robert Matics, Sandor Varga, Balazs Opper, 

Akos Klein, Gyozo Horvath, Alexandre Roulin, Peter Putnoky, and Gyula Hoffmann 142 

Breeding Biology and Food Habits of the Madagascar Kestrel {Falco newtoni) 

IN Northeastern Madagascar. Lily-Arison Rene de Roland, Jeanneney Rabearivony, 

Harilalaina Robenarimangason, Gilbert Razafimanjato, and Russell Thorstrom 1 49 

Short Communications 

Interspecific Aggression and Nest-site Competition in a European Owl Community. 

Inigo Zuberogoitia, Jose Antonio Martinez, Jabi Zabala, and Jose Enrique Martinez 156 

Prey Partitioning between Mates in Breeding Booted Eagles (Hieraaetus pennatus) . 

Jose E. Martinez and Jose F. Calvo 159 

Predation of Small Mammals by Rufous-legged Owl, Barn Owl, and Magellanic Horned Owl 
IN Argentinean Patagonia Forests. Daniel E. Udrizar Sauthier, Analia Andrade, and 
Ulyses F. J. Pardinas 163 

Changes in Site Occupancy and Nesting Performance of Peregrine Falcons in Colorado, 

1963-2004. James H. Enderson 166 

Analysis of Reservoir Selection by Wintering Ospreys (Pandion haliaetus hauaetus) in Andalusia, Spain: 

A Potential Tool for Reintroduction. Eva Casado and Miguel Ferrer 168 

Introduced Animals in the Diets of the Ogasawara Buzzard, an Endemic Insular Raptor in 

the Pacific Ocean. Yuka Kato and Tadashi Suzuki 173 


The Diet of Eurasian Griffons {Gyps fulvus) in Crete. Stavros M. Xirouchakis 


179 


Letters 

Are Earlier Estimates of Accipitriformes Crossing the Channel of Sicily (Central Mediterranean) 

During Spring Migration Accurate? Nicolantonio Agostini 184 

Ground Nesting by Egyptian Vultures {Neophron percnoptekus) in the Canary Islands. 

Laura Gangoso and Cesar-Javier Palacios 186 

First Summer Records of Ospreys {Pandion haliaetus) Along the Coast of Oaxaca, Mexico. 

Juan Meraz and Betzabeth Gonzalez-Bravo 187 


Number 3 

Introduction 

Preface: Proceedings of the International Symposium on the Ecology and Management of 

Northern Goshawks. Clint W. Boal 189 

In Memoriam: Suzanne Merideth Joy. Richard T. Reynolds 190 

Status 

Technical Review of the Status of Northern Goshawks in the Western United States. 

David E. Andersen, Stephen DeStefano, Michael I. Goldstein, Kimberly Titus, Cole 
Crocker-Bedford, John J. Keane, Robert G. Anthony, and Robert N. Rosenfield 192 

Biology 

Is Fledging Success a Reliable Index of Fitness in Northern Goshawks? 

J. David Wiens and Richard T Reynolds 210 

Productivity and Mortai.ity of Northern Goshawks in Minnesota. 

Clint W. Boal, David E. Andersen, and Patricia L. Kennedy 222 

Relationships Between Winter and Spring Weather and Norihern Goshawk {AcapiTER gentilis) 

Reproduction in Northern Nevada. Graham D. Fairhurst and Marc J. Bechard 229 

Patterns of Temporal Variation in Goshawk Reproduction and Prey Resources. 

Susan R. Salafsky, Richard T. Reynolds, and Barry R. Noon 237 

A Skewf.d Sex Ratio in Northern Goshawks; Is It a Sign of a Stressed Population? 

Michael E Ingraldi 247 

Northern Goshawk {Accjpiter gentilis laingi) Post-fledging Areas on Vancouver Isiand, British 

Columbia. Erica L. McClaren, Patricia L. Kennedy, and Donald D. Doyle 253 

Northern Goshawk Diet in Minnesota: An Anaiasis Using Video Recording 

Systems. Brett L. Smithers, Clint W. Boal, and David A. Andersen 264 

Techniques 

Sampling Considerations for Demographic and Habitat Studies of Northern Goshawks. 

Richard T. Reynolds, J. David Wiens, Suzanne M. Joy, and Susan R. Salafsky 274 

Population Genetics and Genotyping for Mark-Recapture Studies of Northern Goshawks 
{Acciitper gentilis) on the Kaibab Plateau, Arizona. Shelley Bayard de Volo, Richard T. 

Reynolds, J, Rick Topinka, Bernie May, and Michael F. Antolin 286 

When Are Goshawks Not There? Is a Single Visit Enough to Infer Absence at Occupied Nest 
Areas? Douglas A. Boyce, Jr., Patricia L. Kennedy, Paul Beier, Michael F. Ingraldi, 

Susie R. MacVean, Melissa S. Siders, John R. Squires, and Brian Woodbridge 296 

Quantifying Northern Goshawk Diets Using Remote Cameras and Observations 

from Blinds. Andi S. Rogers, Stephen DeStefano, and Micheal F. Ingraldi 303 


Conservation 


Temporal Pati'erns of Northern Goshawk Nest Area Occupancy and Habitat: A Retrospective 

Anai.ysis. Steven Desimone and Stephen DeStefano 310 

Monitoring Results of Northern Goshawk Nesting Areas in the Greater Yellowstone Ecosystem: 

Is Decline in Occupancy Reiated to Habitat Change? Susan M. Patla 324 

Effects of Timber Harvesting Near Nest Sites on the Reproductive Success of Northern 

Goshawks {Accipiter gentius). Todd Mahon and Frank I. Doyle 335 

A Review of the Status and Distribution of Northern Goshawks in New England. 

Stephen DeStefano 342 


Number 4 


Taxonomic Status and Biology of the Cuban Black-Hawk, Buteogallus 

ANTHRACINUS GUNDLACHII (AVES: AcCIPITRIDAE) . James W. Wiley and Orlando H. Garrido.... 351 


Home Range and Habitat Use of Northern Spotted Owls on the Olympic 

Peninsula, Washington. Eric D. Forsman, TimmothyJ. Kaminski, Jeffery C. Lewis, 

Kevin J. Maurice, Stan G. Sovern, Cheron Ferland, and Elizabeth M. Glenn 365 

First-cycle Molts in North American Falconiformes. Peter Pyle 378 

Morphometric Analysis of Large Falco Species and Their Hybrids with 
Implications for Conservation. Chris p. Eastham and Mike k. Nichoiis 386 

A Change in Foraging Success and Cooperative Hunting by a Breeding Pair 

OF Peregrine Falcons and Their Fledglings. Dick Dekker and Robert Taylor 394 

Nesting Ecology and Behavior of Broad-winged Hawks in Moist Karst 

Forests of Puerto Rico. Derek W. Hengstenberg and Francisco J. Vilella 404 


Raptor Abundance and Distribution in the Llanos Wetlands of Venezuela. 

Wendy J. Jensen, Mark S. Gregory, Guy A. Baldassarre, Francisco J. Vilella, and Keith L. Bildstein .... 417 


A Comparison of Breeding Season Food Habits of Burrowing Owls Nesting 
IN Agricultural and Nonagricultural Habitat in Idaho. Colleen e. Moulton, 

Ryan S. Brady, and James R. Belthoff 429 

Red-tailed Hawk Dietary Overlap with Northern Goshawks on the Kaibab 

Plateau, Arizona. Angela E. Gatto, Teryl G. Grubb, and Carol L. Chambers 439 

Bat Predation by Long-eared Owls in Mediterranean and Temperate Regions 

OF Southern Europe. Ana Marfa Garcfa, Francisco Cervera, and Alejandro Rodriguez 445 

Short Communications 

Differential Effectiveness of Playbacks for Little Owls {Athene noctua) Surveys before and 
AFTER Sunset. Joan Navarro, Eduardo Mfnguez, David Garcfa, Carlos Villacorta, Francisco 
Botella, JosJ Antonio S<nchez-Zapata, Martina Carrete, and Andijs Gimjnez 454 


King Vultures {Sarcoramphus papa) Forage in Moriche and Cucurit Palm Stands. Marsha A. Schlee 458 


Family Break Up, Departure, and Autumn Migration in Europe of a Family of Greater Spotted 
Eagles (Aquila clanga) as Reported by Satellite Telemetry. Bernd-U. Meyburg, Christiane 
Meyburg, Tadeusz Mizera, Grzegorz Maciorowski, and Jan Kowalski 462 

Seasonal Patterns of Common Buzzard {Buteo buieo) Relative Abundance and Behavior in 

POLLINO National Park, Italy. Massimo Pandolfi, Alessandro Tanferna, and Giorgia Gaibani 466 

New Nesting Record and Observations of Breeding Peregrine Falcons in Baja California Sur, 

Mexico. Aradit Castellanos, Cerafma Argiielles, Federico Salinas-Zavala, and Alfr edo Ortega-Rubio 472 

Letters 

A Previously Undescribed Vocalization of the Northern Pygmv-Owl. Graham G. Frye 476 

Book Review. Edited by Jeffrey S. Marks 478 

Information for Contributors 480 

Index to Volume 39 484 


A Telemetry Receiver Designed with 
The Researcher in Mind 


What you've been waiting fon 



finally, a highly s^n\itiv« 999 channel iynthesi/cd telemetry receiver that weiqhs less than 13 ourtcts. Is 
completely user programmable and offers variable scan rates over all frequencies. For each animal being 
tracked, the large LCD display provides not only the frequency (to lOOHz) and channel number, but also a 
7 character alphanumenc comment held and a digital signal strength meter. Stop carrying receivers that are 
the si/e of a lunch bo* or cost over SlSOO. The features and performance of the ne»v R>1000 pocket sized 
telemetry receiver will impress you. and the prne will convince you. 

Othtr ftatum Imludt: 

• factory toned to any 4MH/ wide segment in the 148 174MHz Band • Very high sensitivity of •148dBm to 
ISOdBm • Illuminated display and keypad for use in low light or darkness • User selectable Kan rates 
from 1-30 seconds in 1 second steps • Rechargeable battenes operate the receiver for 12 hours and 

can be replaced with standard AA Alkaline batteries in the field. 
Both 12vdc and llOvac chargers are included. 

• b.r (IS.bcm) high, 

2.6" (6.6cm) wide. 

1.5" (3.8cm) deep. 

• 3 year warranty 

• 1 day delivery 

$ 695.00 

Please specify desired 4MHz 
wide segment in the 
148-174HHZ band 

Visit our 
website for 
complete 
specifications, 
operating 
manual and 
information 
on the R- 1000 
or call our 
toll free number 
to order your 
receiver now. 

Try th€ 

New R IOOO 
and YduH 
Impressed! 




COMMUNICATIONS SPECIALISTS. INC. 

■ • .''ri ('i-- ! A.I-' ! ; 1 A OCj • 1 




'^.^0 


r, •.! 


.S .S A 



Buteq 
Books 


Toll Free: 800-722-2460 
phone: 434-263-8671 
fax: 434-263-4842 

Specializing in Ornithology 


Buteo Books is the largest retailer of Ornithology books in North America, 
with over 2,000 in-print titles, and hundreds of out-of-print titles available. 



Rffutyt 

P^m 



NEW, IN-PRINT ORNITHOLOGY 

The most popular authors, cutting edge studies, and unusual subjects. We stock scientific 
texts, hard-to-find foreign books, and entertaining reading; plus audio, video, and software. 

RARE AND OUT-OF-PRINT ORNITHOLOGY 

Buteo Books has a wide selection of used, out-of-print, and scarce titles available, including 
classics on birds of prey and falconry. Our stock changes weekly, so call to check availability. 


BIRDS OF NORTH AMERICA SERIES 

This series consists of individual species accounts for each of the 
species which breed in the United States and Canada, including 
diurnal and nocturnal raptors. These illustrated reviews provide 
comprehensive summaries of the current knowledge of each species, 
with range maps and extensive lists of references. 

As the exclusive distributor of print copies in this series, Buteo Books 
is pleased to offer these accounts for $7.50 each. All 716 profiles are 
listed in taxonomic order on our website. 


■a 


Buteo Books; 3130 Laurel Road; Shipman, VA 22971; USA 


Visit our website for more information; 

www.buteobooks.com 



r Essential Guides 



"One of the best bird field guides ever published." 

— Oriental Bird Club Bulletin 

Birds of Southeast Asia 

Craig Robson 

• More than 140 full-color plates 

• All 1 ,270 species covered in detail 

• Up-to-date text covers identification, voice, habitat, behavior, 
and range of all the region's species and distinctive subspecies 

• Complete coverage of some fifteen Southeast Asian countries 
and regions 

304 pages. 142 color plates. 5 3/4 x 8 1/4. 

Princeton Field Guides Series 
Paper $29.95 0-691-12435-3 

Not for sale in the Commonwealth (except Canada) and the European Union 


"Jerry Liguori has created a book that can easily be 
toted in the field, and is an absolute must-have for 
any raptor enthusiast!"— Brian K. Wheeler, author of 
Raptors of Eastern/Wesfern North America 

Hawks from Every Angle 

How to Identify Raptors In Flight 

Jerry Liguori 

Foreword by David Allen Sibley 

• The essential new approach to identifying hawks in flight 

• Innovative, accurate, and field-tested 

• Compares and contrasts species easily confused with 
one another 

• Provides the pictures (and words) needed for identifica- 
tion in the field 

• Covers the 1 9 species common to North America 

• User-friendly format for quick use 

339 color photos. 32 b/w photos. 7 1/2x9 1/2. 

Paper $19.95 0-691-11825-6 Cloth $55.00 0-691-11824-8 



Cclchratint^ lOO Years of Excellence 

PRINCETON 






.1 , ’ ■ Kf jd cxccrpls or-inc 

‘Umversitv ‘.Press 



V 



2006 ANNUAL MEETING 


The Raptor Research Foundation, Inc. 2006 annual meeting will be held in conjunction with the 
Fourth North American Ornithological Conference on 3-7 October 2006 in Veracruz, Mexico. For 
more information about the meeting see http:/ /www.naoc2006.org/ 


Persons interested in predatory birds are invited to join The Raptor Research Foundation, Inc. (see: 
http:/ /biology.boisestate.edu/ raptor/). Send requests for information concerning membership, subscriptions, 
special publications, or change of address to OSNA, 5400 Bosque Blvd., Suite 680, Waco TX 76710, U.S.A. 

The, Journal of Raptor Research (ISSN 0892-1016) is published quarterly and available to individuals for $40.00 
per year and to libraries and institutions for $65.00 per year from The Raptor Research Foundation, Inc., 14377 
1 17th Street South, Hastings, Minnesota 55033, U.S.A. (Add $3 for destinations outside of the continental United 
States.) Periodicals postage paid at Hastings, Minnesota, and additional mailing offices. POSTMASTER: Send 
address changes to The Journal of Raptor Research, OSNA, P.O. Box 1897, Lawrence, KS 66044-8897, U.S.A. 

Printed by Allen Press, Inc., Lawrence, Kansas, U.S.A. 

Copyright 2005 by The Raptor Research Foundation, Inc. Printed in U.S.A. 

@ This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). 


Raptor Research Foundation, Inc. 

Grants and Awards 

For details and additional information visit: 
http:/ /biology.boisestate.edu/raptor/grants%20and%20awards.htm 

Awards for Recognition of Significant Contributions 

The Tom Cade Award is a non-monetary award that recognizes an individual who has made significant advances 
in the area of captive propagation and reintroduction of raptors. The Fran and Frederick Hamerstrom 
Award is a non-monetary award that recognizes an individual who has contributed significantly to the under- 
standing of raptor ecology and natural history. Submit nominations for either award to: Dr. Clint Boal, Texas 
Cooperative Fish and Wildlife Research Unit, BRD/USGS, Texas Tech University, 15th Street & Boston, 
Ag Science Bldg., Room 218, Lubbock TX 79409-2120 U.S.A.; phone: 806-742-2851; e-mail: cboal@ttu.edu 

Awards for Student Recognition and Travel Assistance 

The James R. Koplin Travel Award is given to a student who is the senior author and presenter of a paper or 
poster to be presented at the RRF meeting for which travel funds are requested. Application deadline: due 
date for meeting abstract. Contact: Dr. Patricia A. Hall, 5937 E. Abbey Rd., Flagstaff, AZ 86004; phone: 
520-526-6222 U.S.A.; e-mail: pah@spruce.for.nau.edu 

The William C. Anderson Memorial Award is given to both the best student oral and poster presentation at the 
annual RRF meeting. The paper cannot be part of an organized symposium to be considered. Application 
deadline; due date for meeting abstract, no special application is needed. Contact: Rick Gerhardt, Sage 
Science, 319 SE Woodside Ct., Madras, OR 97741 U.S.A; phone: 541-475-4330; email; rgerhardt@madras.net 

Grants 

Application deadline for all grants is February 15 of each year; selections will be made by April 15. 

The Dean Amadon Grant for up to $1000 is designed to assist persons working in the area of systematics (tax- 
onomy) and distribution of raptors. The Stephen R. TuUy Memorial Grant for up to $500 is given to sup- 
port research and conservation of raptors, especially to students and amateurs with limited access to alter- 
native funding. Agency proposals are not accepted. Contact for both grants; Dr. Carole Griffiths, 251 
Mar ding Ave., Tarrytown, NY 10591 U.S.A.; phone: 914-631-2911; e-mail; cgriff@liu.edu 

The Leslie Brown Memorial Grant for up to $1400 is given to support research and/or the dissemination of 
information on African raptors. Contact: Dr. Jeffrey L. Lincer, 9251 Golondrina Drive, La Mesa, CA 91941, 
U.S.A.; e-mail: JeflLincer@tns.net